0707.0839/pgw.tex
1: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2: %
3: %   Juan Garcia-Bellido, Daniel G. Figueroa and Alfonso Sastre
4: %
5: %   A Gravitational Wave Background from Reheating after Hybrid Inflation
6: %
7: %   Version 1.12   November 26, 2007
8: %
9: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
10: 
11: \documentclass[prd,twocolumn,aps,amsmath,amssymb,nofootinbib,preprintnumbers,fleqn]
12: {revtex4}
13: \special{papersize=8.5in,11in}
14: \voffset=1cm
15: \usepackage{graphicx}% Include figure files
16: \usepackage{dcolumn}% Align table columns on decimal point
17: \usepackage{bm}% bold math
18: \usepackage{amsmath}
19: \usepackage{amsfonts}
20: \usepackage{subfigure}
21: % Young tableaux
22: % draw box with width #1pt and line thickness #2pt
23: \newcommand{\drawsquare}[2]{\hbox{%
24: \rule{#2pt}{#1pt}\hskip-#2pt%  left vertical
25: \rule{#1pt}{#2pt}\hskip-#1pt%  lower horizontal
26: \rule[#1pt]{#1pt}{#2pt}}\rule[#1pt]{#2pt}{#2pt}\hskip-#2pt%upper horizontal
27: \rule{#2pt}{#1pt}}% right vertical
28: 
29: \newcommand{\Yfund}{\raisebox{-.5pt}{\drawsquare{6.5}{0.4}}}%  fund
30: \newcommand{\Yasymm}{\raisebox{-3.5pt}{\drawsquare{6.5}{0.4}}\hskip-6.9pt%
31:                      \raisebox{3pt}{\drawsquare{6.5}{0.4}}%
32:                     }%  antisymmetric second rank
33: \newcommand{\Ysymm}{\Yfund\hskip-0.4pt%
34:                     \Yfund}%  symmetric second rank
35: \def\symm{\Ysymm}
36: \def\bsymm{\overline{\Ysymm}}
37: \def\ls{\mathrel{\lower4pt\vbox{\lineskip=0pt\baselineskip=0pt
38:            \hbox{$<$}\hbox{$\sim$}}}}
39: \def\gs{\mathrel{\lower4pt\vbox{\lineskip=0pt\baselineskip=0pt
40:            \hbox{$>$}\hbox{$\sim$}}}}
41: %  draw box of size #1pt and line thickness #2pt
42: \def\drawbox#1#2{\hrule height#2pt
43: 
44: \hbox{\vrule width#2pt height#1pt \kern#1pt
45:               \vrule width#2pt}
46:               \hrule height#2pt}
47: 
48: \def\Fund#1#2{\vcenter{\vbox{\drawbox{#1}{#2}}}}
49: \def\Asym#1#2{\vcenter{\vbox{\drawbox{#1}{#2}
50:               \kern-#2pt       % line up boxes
51:               \drawbox{#1}{#2}}}}
52: \def\sym#1#2{\vcenter{\hbox{ \drawbox{#1}{#2} \drawbox{#1}{#2}    }}}
53: \def\fund{\Fund{6.4}{0.3}}
54: \def\asymm{\Asym{6.4}{0.3}}
55: \def\bfund{\overline{\fund}}
56: \def\basymm{\overline{\asymm}}
57: 
58: %%%%% end Yang
59: 
60: \def\half{\frac{1}{2}}
61: 
62: \newcommand{\beq}{\begin{equation}}
63: \newcommand{\eeq}{\end{equation}}
64: 
65: \begin{document}
66: 
67: \title{A Gravitational Wave Background from Reheating after Hybrid Inflation}
68: 
69: \author{Juan Garc\'\i a-Bellido, Daniel G. Figueroa and Alfonso Sastre}
70: \affiliation{Departamento de F\'\i sica Te\'orica, 
71: Universidad Aut\'onoma de Madrid, Cantoblanco, 28049 Madrid, Spain\\
72: Instituto de F\'\i sica Te\'orica CSIC-UAM, 
73: Universidad Aut\'onoma de Madrid, Cantoblanco, 28049 Madrid, Spain}
74: 
75: 
76: \date{November 26th, 2007}
77: 
78: \begin{abstract}
79: The reheating of the universe after hybrid inflation proceeds through
80: the nucleation and subsequent collision of large concentrations of
81: energy density in the form of bubble-like structures moving at
82: relativistic speeds. This generates a significant fraction of energy
83: in the form of a stochastic background of gravitational waves, whose
84: time evolution is determined by the successive stages of
85: reheating: First, tachyonic preheating makes the amplitude of gravity
86: waves grow exponentially fast. Second, bubble collisions add a new
87: burst of gravitational radiation. 
88: Third, turbulent motions finally sets the end of gravitational waves 
89: production. From then on, these waves propagate unimpeded to us.
90: We find that the fraction of energy density today in these
91: primordial gravitational waves could be significant for GUT-scale
92: models of inflation, although well beyond the frequency range
93: sensitivity of gravitational wave observatories like LIGO, LISA or
94: BBO. However, low-scale models could still produce a detectable signal
95: at frequencies accessible to BBO or DECIGO. For comparison, we have
96: also computed the analogous gravitational wave background from some 
97: chaotic inflation models and obtained
98: results similar to those found by other groups. The discovery of such
99: a background would open a new observational window into the very early
100: universe, where the details of the process of reheating, i.e. the Big
101: Bang, could be explored.  Moreover, it could also serve in the future
102: as a new experimental tool for testing the Inflationary Paradigm.
103: \end{abstract}
104: \preprint{IFT-UAM/CSIC-07-38}
105: \maketitle
106: 
107: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
108: 
109: 
110: \section{Introduction}
111: 
112: Gravitational waves (GW) are ripples in space-time that travel at the
113: speed of light, and whose emission by relativistic bodies represents a
114: robust prediction of General Relativity.  The change in the orbital
115: period of a binary pulsar known as PSR 1913+16 was used by Hulse and
116: Taylor~\cite{HulseTaylor} to obtain indirect evidence of their
117: existence. Although gravitational radiation has not been directly
118: detected yet, it is expected that the present universe should be
119: permeated by a diffuse background of GW of either an astrophysical or
120: cosmological origin~\cite{Maggiore}. Astrophysical sources, like the
121: gravitational collapse of supernovae or the neutron star and black
122: hole binaries' coalescence, produce a stochastic gravitational wave
123: background (GWB) which can be understood as coming from unresolved
124: point sources. On the other hand, among the backgrounds of
125: cosmological origin, we find the approximately scale-invariant
126: background produced during inflation~\cite{Starobinsky}, or the GWB
127: generated at hypothetical early universe thermal phase transitions,
128: from relativistic motions of turbulent plasmas or from the decay of
129: cosmic strings~\cite{Maggiore}. Fortunately, these backgrounds have
130: very different spectral shapes and amplitudes that might, in the
131: future, allow gravitational wave observatories like the Laser
132: Interferometer Gravitational Wave Observatory (LIGO)~\cite{LIGO}, the
133: Laser Interferometer Space Antenna (LISA)~\cite{LISA}, the Big Bang
134: Observer (BBO)~\cite{BBO} or the Decihertz Interferometer
135: Gravitational Wave Observatory (DECIGO)~\cite{DECIGO}, to disentangle
136: their origin~\cite{Maggiore}. Unfortunately, due to the weakness of
137: gravity, this task will be extremely difficult, requiring a very high
138: accuracy in order to distinguish one background from another. It is
139: thus important to characterize as many different sources of GW as
140: possible.
141: 
142: There are, indeed, a series of constraints on some of these
143: backgrounds, the most stringent one coming from the large-scale
144: polarization anisotropies in the Cosmic Microwave Background (CMB),
145: which may soon be measured by Planck~\cite{Planck}, if the scale of
146: inflation is sufficiently high. There are also constraints coming from
147: Big Bang nucleosynthesis~\cite{BBN}, since such a background would
148: contribute as a relativistic species to the expansion of the universe
149: and thus increase the light element abundance. There is also a
150: constraint coming from millisecond pulsar timing~\cite{pulsar}.
151: Furthermore, it has recently been proposed a new constraint on a GWB
152: coming from CMB anisotropies~\cite{Elena}. Most of these constraints
153: come at very low frequencies, typically from $10^{-18}$ Hz to
154: $10^{-8}$ Hz, while present GW detectors work at frequencies of order
155: 1-100 Hz, and planned observatories will range from $10^{-3}$ Hz of
156: LISA to $10^{3}$ Hz of Advanced-LIGO~\cite{Maggiore,LIGO}. If early
157: universe phenomena like first order phase
158: transitions~\cite{KosowskyTurner,Nicolis} or cosmic
159: turbulence~\cite{Turbulence} occurred around the electro-weak (EW)
160: scale, there is a chance than the GW detectors will measure the
161: corresponding associated backgrounds. However, if such early universe
162: processes occurred at the GUT scale, their corresponding backgrounds
163: will go undetected by the actual detectors, since these cannot reach
164: the required sensitivity in the high frequency range of $10^7-10^9$
165: Hz, corresponding to the size of the causal horizon at that time.
166: There are however recent proposals to cover this
167: range~\cite{Nishizawa2007}, which may become competitive in the near
168: future.
169: 
170: Moreover, present observations of the CMB anisotropies and
171: the Large Scale Structure (LSS) distribution of matter seem to suggest
172: that something like Inflation must have occurred in the very early
173: universe. We ignore what drove inflation and at what scale it took
174: place. However, approximately scale-invariant density perturbations,
175: sourced by quantum fluctuations during inflation, seem to be the most
176: satisfying explanation for the CMB anisotropies. Together with such
177: scalar perturbations one also expects tensor perturbations (GW) to be
178: produced, with an almost scale-free power
179: spectra~\cite{Starobinsky}. Because of the weakness of gravity, this
180: primordial inflationary GWB should decouple from the rest of matter as
181: soon as it is produced, and move freely through the Universe till
182: today. At present, the biggest efforts employed in the search for
183: these primordial GW come from the indirect effect that this
184: background has on the B-mode polarization anisotropies of the
185: CMB~\cite{Planck}, rather than via direct detection. The detection of
186: such a background is crucial for early universe cosmology because it
187: would help to determine the absolute energy scale of inflation, a
188: quantity that for the moment is still uncertain, and would open the
189: exploration of physics at very high energies.
190: 
191: In the early universe, after inflation, other backgrounds of GW could
192: have been produced at shorter wavelengths, in a more 'classical'
193: manner, rather than sourced by quantum fluctuations. In particular,
194: whenever there are large and fast moving inhomogeneities in the matter
195: distribution, one expects the emission of GW.  This is much like the
196: situation in classical electrodynamics, but with some differences. At
197: large distances from the source, the amplitude of the electromagnetic
198: field $A_i$ is expressed as the first derivative of the dipole moment
199: $d_i$ of the charge distribution of the source, $A_i\simeq \dot d_i /
200: cr$, while the amplitude of the GW is given by the second derivative
201: of the quadrupole moment of the mass distribution, $h_{ij} \simeq
202: G\ddot Q_{ij} / c^4r$. In both cases, the larger the velocity of the
203: matter/charge distribution, the larger the amplitude of the radiation
204: produced. Nevertheless, the main difference between the two cases is
205: the weakness of the strength of gravity to that of
206: electromagnetism. Thus, in order to produce a significant amount of
207: gravitational radiation, it is required that the motion of huge masses
208: occurs at speeds close to that of light for the case of astrophysical
209: sources, or a very fast motion and high density contrasts in the
210: continuous matter distribution for the case of cosmological sources.
211: In fact, this is believed to be the situation at the end of inflaton,
212: during the conversion of the huge energy density driving inflation
213: into radiation and matter at the so-called {\it reheating} of the
214: Universe~\cite{preheating}. Such an event corresponds to 
215: the actual Big Bang of the Standard Cosmological Model.
216: 
217: Note that any background of GW coming from the early universe, if
218: generated below Planck scale, immediately decoupled upon production,
219: as can be easily understood by the following dimensional analysis
220: argument. Assuming that gravitons were in thermal equilibrium with the
221: early universe plasma, at a temperature $T$, the gravitons' cross
222: section should be of order $\sigma\sim G^2T^2$. Then, given the
223: graviton number density $n\sim T^3$ and velocity $v = 1$, the
224: gravitons' interaction rate should be $\Gamma = \left\langle n\sigma
225: v\right\rangle\,\sim\,T^5/M_p^4$. Since the Hubble rate is $H\sim
226: T^2/M_p$, then $\Gamma\,\sim\,H\,(T/M_p)^3$, so gravitons could {\it
227: not} be kept in equilibrium with the surrounding plasma for
228: $T\,<\,M_p$. Therefore, GW produced well after Planck scale will
229: always be decoupled from the plasma, and whatever their spectral
230: signatures, they will retain their shape throughout the expansion of
231: the Universe. Thus, the characteristic frequency and shape of the GWB
232: generated at a given time should contain information about the very
233: early state of the Universe in which they were produced.  Actually, it
234: is conceivable that, in the not so far future, the detection of these
235: GW backgrounds could be the only way we may have to infer the physical
236: conditions of the Universe at such high energy scales, which certainly
237: no particle collider will ever reach. However, the same reason that
238: makes GW ideal probes of the early universe $-$ the weakness of gravity
239: $-$, is responsible for the extreme difficulties we have for their
240: detection on Earth. For an extensive discussion see Ref.~\cite{Thorne}.
241: 
242: In a recent letter~\cite{GBF} we described the stochastic background
243: predicted to arise from reheating after hybrid inflation. In this
244: paper we study in detail the various processes involved in
245: the production of such a background, whose detection could open a new
246: window into the very early universe. In the future, this background
247: could also serve as a new tool to discriminate among different
248: inflationary models, as each of these would give rise to a different
249: GWB with very characteristic spectral features. The first stage of the
250: energy conversion at the end of inflation,
251: preheating~\cite{preheating}, is known to be explosive and extremely
252: violent, and quite often generates in less than a Hubble time the huge
253: entropy measured today. The details of the dynamics of preheating
254: depend very much on the model and are often very complicated because
255: of the non-linear, non-perturbative and out-of-equilibrium character
256: of the process itself. However, all the cases have in common that only
257: specific resonance bands of the fields suffer an exponential
258: instability, which makes their occupation numbers grow by many orders
259: of magnitude. The shape and size of the spectral bands depend very
260: much on the inflationary model. If one translates this picture into
261: position-space, the highly populated modes correspond to large
262: time-dependent inhomogeneities in the matter distributions which
263: acts, in fact, as the source of GW we are looking for.
264: 
265: For example, in single field chaotic inflation models, the coherent
266: oscillations of the inflaton during preheating generates, via
267: parametric resonance, a population of highly occupied modes that
268: behave like waves of matter. They collide among themselves and their
269: scattering leads to homogenization and local thermal equilibrium.
270: These collisions occur in a highly relativistic and very asymmetric
271: way, being responsible for the generation of a stochastic
272: GWB~\cite{TkachevGW,JuanGW,EastherLim} with a typical frequency today
273: of the order of $10^{7} - 10^{9}$ Hz, corresponding to the present
274: size of the causal horizon at the end of high-scale inflation. There
275: is at present a couple of experiments searching for such a background,
276: see Refs.~\cite{Nishizawa2007}, based of laser interferometry, as well
277: as by resonant superconducting microwave cavities~\cite{Picasso}.
278: 
279: However, there are models like hybrid inflation in which the end of
280: inflation is sudden~\cite{hybrid} and the conversion into radiation
281: occurs almost instantaneously. Indeed, since the work of
282: Ref.~\cite{tachyonic} we know that hybrid models preheat in an even
283: more violent way than chaotic inflation models, via the spinodal
284: instability of the symmetry breaking field that triggers the end of
285: inflation, irrespective of the couplings that this field may have to
286: the rest of matter. Such a process is known as \textit{tachyonic
287: preheating}~\cite{tachyonic,symmbreak} and could be responsible for
288: copious production of dark matter particles~\cite{ester}, lepto and
289: baryogenesis~\cite{CEWB}, topological defects~\cite{tachyonic},
290: primordial magnetic fields~\cite{magnetic}, etc.
291: 
292: It was speculated in Ref.~\cite{JuanGW} that in (low-scale) models of
293: hybrid inflation it might be possible to generate a stochastic GWB in
294: the frequency range accessible to present detectors, if the scale of
295: inflation was as low as $H_{\rm inf} \sim 1$ TeV. However, the
296: amplitude was estimated using the parametric resonance formalism of
297: chaotic preheating, which may not be applicable in this case. In
298: Ref.~\cite{symmbreak} (from now on referred to as paper I), it was
299: shown that the process of symmetry breaking proceeds via the
300: nucleation of dense bubble-like structures moving at the speed of
301: light, which collide and break up into smaller structures (see Figs.~7
302: and~8 of paper I). We conjectured at that time that such collisions
303: would be a very strong source of GW, analogous to the GW production
304: associated with strongly first order phase
305: transitions~\cite{KosowskyTurner}. As we will show in this paper, this
306: is indeed the case during the nucleation, collision and subsequent
307: rescattering of the initial bubble-like structures produced after
308: hybrid inflation. During the different stages of reheating in this
309: model, gravity waves are generated and amplified until the Universe
310: finally thermalizes and enters into the initial radiation era of the
311: Standard Model of Cosmology.  From that moment until now, during the
312: whole thermal history of the expansion of the universe, this cosmic
313: GWB will be redshifted as a radiation-like fluid, totally decoupled
314: from any other energy-matter content of the universe, such that
315: today's ratio of energy stored in these GW to that in radiation, could
316: range from $\Omega_{_{\rm GW}}h^2\sim 10^{-8}$, peacked around $f\sim
317: 10^7$~Hz for the high-scale models, to $\Omega_{_{\rm GW}}h^2\sim
318: 10^{-11}$, peacked around $f\sim 1$ Hz for the low-scale models.
319: 
320: Finally, let us mention that since the first paper by Khlebnikov and
321: Tkachev ~\cite{TkachevGW}, studing the GWB produced at reheating after
322: chaotic inflation, it seems appropriate to reanalyze this topic in a
323: more detailed way. The idea was extended to hybrid inflation
324: in~\cite{JuanGW}, but within the parametric resonance formalism. It was
325: also revisited very recently in Ref.~\cite{EastherLim,EastherLim2} for
326: the $\lambda\phi^4$ and $m^2\phi^2$ chaotic scenarios, and reanalysed
327: again for hybrid inflation in Ref.~\cite{GBF}, this time using the new
328: formalism of tachyonic preheating~\cite{tachyonic,symmbreak}.  Because
329: of the increase in computer power of the last few years, we are now
330: able to perform precise simulations of the reheating process in a
331: reasonable time scale. Moreover, understanding of reheating has
332: improved, while gravitational waves detectors are beginning to attain
333: the aimed sensitivity~\cite{LIGO}. Furthermore, since these cosmic
334: GWBs could serve as a deep probe into the very early universe, we
335: should characterize in the most detailed way the information that we
336: will be able to extract from them.
337: 
338: The paper is divided as follows. In Section II we briefly review the
339: hybrid model of inflation. Section III is dedicated to our approach
340: for extracting the power spectrum of GW from reheating. In section
341: IV, we give a detailed account of the lattice simulations performed
342: with two codes: our own FORTRAN parallelized computer code (running in
343: the MareNostrum supercomputer~\cite{BSC} and in our UAM-IFT
344: cluster~\cite{IFTcluster}), as well as with a modified version of the
345: publicly available C++ package LATTICEEASY \cite{LatticeEasy}. Section
346: V is dedicated to study the spatial distribution of the
347: production of gravitational waves. In Section VI, we reproduce as a
348: crosscheck, some of the results of~\cite{TkachevGW,EastherLim,DufauxGW} 
349: concerning the GWB produced at reheating after chaotic inflation models. 
350: Finally, in section VII, we give our conclusions and perspectives for 
351: the future.
352: 
353: \section{The Hybrid Model}
354: 
355: Hybrid inflation models~\cite{hybrid} arise in theories of particle
356: physics with symmetry breaking fields ('Higgses') coupled to flat
357: directions, and are present in many extensions of the Standard Model,
358: both in string theory and in supersymmetric theories~\cite{LythRep}. 
359: The potential in these models is given by
360: \begin{equation}\label{higgsPotential}
361: \indent V(\Phi,\chi) = \lambda\left(\Phi^{\dagger}\Phi - 
362: {v^2\over2}\right)^2 + g^2\chi^2\Phi^{\dagger}\Phi 
363: + \frac{1}{2}\mu^2\chi^2\,,
364: \end{equation}
365: where the contraction $\Phi^{\dagger}\Phi$ should be understood as the
366: trace ${\rm Tr}\,\Phi^{\dagger}\Phi=\half|\phi|^2$.  Inflation occurs
367: along the lifted flat direction, satisfying the slow-roll conditions
368: thanks to a large vacuum energy $\rho_0=\lambda v^4/4$. Inflation ends
369: when the inflaton $\chi$ falls below a critical value and the symmetry
370: breaking field $\phi$ acquires a negative mass squared, which triggers
371: the breaking of the symmetry and ends in the true vacuum, $|\phi|=v$,
372: within a Hubble time. These models do not require small couplings in
373: order to generate the observed CMB anisotropies; e.g. a working model
374: with GUT scale symmetry breaking, $v=10^{-3}\,M_P$, with a Higgs
375: self-coupling $\lambda$ and a Higgs-inflaton coupling $g$ given by
376: $g=\sqrt{2\lambda}=0.05$, satisfies all CMB constraints~\cite{WMAP},
377: and predicts a tiny tensor contribution to the CMB polarization. The
378: main advantage of hybrid models is that, while most chaotic inflation
379: models can only occur at high scales, with Planck scale values for the
380: inflaton, and $V_{\rm inf}^{1/4} \sim 10^{16}$ GeV, one can choose the
381: scale of inflation in hybrid models to range from GUT scales all the
382: way down to GeV scales, while the Higgs \textit{v.e.v.} can range from
383: Planck scale, $v=M_P$, to the Electroweak scale, $v=246$ GeV, see
384: Ref.~\cite{hybrid,CEWB}.
385: 
386: There are a series of constraints that a hybrid inflation model should
387: satisfy in order to be in agreement with observations. First of all,
388: inflation should end in less than one e-fold, otherwise unacceptable
389: black holes would form~\cite{GBLW}. This can be written as the 
390: \textit{waterfall condition}~\cite{hybrid}, $\lambda m M_P^2 \gg M^3$, 
391: which becomes
392: \begin{eqnarray}
393: \indent {m\over M} \gg {v^2\over M_P^2}\,.
394: \end{eqnarray} 
395: Then there is the condition, known as the COBE normalization, that the
396: scalar amplitude should satisfy $A_S = H^2/2\pi\dot\phi \simeq
397: 5\times10^{-5}$, which gets translated into
398: \begin{eqnarray}
399: \indent g = (n-1){M_P\over v}\sqrt{3\pi\over8}\times10^{-4}\,e^{(n-1)N/2}
400: \end{eqnarray} 
401: as well as the spectral tilt, 
402: \begin{eqnarray}
403: \indent n - 1 = {1\over\pi}{m^2\over M^2}{M_P^2\over v^2} < 0.05
404: \end{eqnarray} 
405: and finally the fact that we have not seen so far any tensor
406: (gravitational wave) contribution in the CMB anisotropies, $r =
407: A_T^2/A_S^2 < 0.3$, imposes the constraint
408: \begin{eqnarray}
409: \indent \lambda^{1/4} < 2\times10^{-3}\,{M_P\over v}\,.
410: \end{eqnarray} 
411: 
412: Taking all these conditions together, we find that a model with $v=0.1
413: M_P$ is probably ruled out, while one with $v=0.01\,M_P$ is perfectly
414: consistent with all observations, and with reasonable values of the
415: coupling constants, e.g. $g=4\times10^{-4}$ and $\lambda = 10^{-3}$.
416: However, the lower is the scale of inflation, the more difficult it is
417: to accommodate the amplitude of the CMB anisotropies with reasonable
418: values of the parameters. For a scale of inflation as low as $10^{11}$
419: GeV, one must significantly finetune the couplings, although there are
420: low scale models based on supersymmetric extensions of the standard
421: model which can provide a good match to observations~\cite{MSSMinf}.
422: 
423: In the following sections we will show how efficient is the production
424: of GW at reheating after hybrid inflation, using both analytical
425: estimates and numerical simulations to derive the amplitude of the
426: present day GWB. Reheating in hybrid inflation~\cite{hybrid} goes
427: through four well defined regimes: first, the exponential growth of
428: long wave modes of the Higgs field via spinodal instability, which
429: drives the explosive growth of all particles coupled to it, from
430: scalars~\cite{tachyonic} to gauge fields~\cite{CEWB} and
431: fermions~\cite{ester}; second, the nucleation and collision of high
432: density contrast and highly relativistic bubble-like structures
433: associated with the peaks of a Gaussian random field like the
434: Higgs, see paper I; third, the turbulent regime that ensues after
435: all these `bubbles' have collided and the energy density in all fields
436: cascades towards high momentum modes; finally, thermalization of all
437: modes when local thermal and chemical equilibrium induces
438: equipartition. The first three stages can be studied in detailed
439: lattice simulations thanks to the semi-classical character of the
440: process of preheating~\cite{classical}, while the last stage is
441: intrinsically quantum and has never been studied in the lattice.
442: 
443: \section{Gravitational Wave Production}
444: 
445: Our main purpose in this paper is to study the details of the
446: stochastic GWB produced during the reheating of the universe after
447: hybrid inflation (sections III, IV and V). However, we also study,
448: albeit very briefly, the analogous background from reheating in some
449: simple chaotic models (section VI). Thus, in this section we derive a
450: general formalism for extracting the GW power spectrum in any scenario
451: of reheating within the (flat) Friedman-Robertson-Walker (FRW)
452: universe. The formalism will be simplified when applied to scenarios
453: in which we can neglect the expansion of the universe, like in the
454: case of most Hybrid models.
455: 
456: A theory with an inflaton scalar field $\chi$ interacting 
457: with other Bose fields $\phi_a$, can be described by
458: \begin{equation}\label{lagrangian}
459: \indent \mathcal{L} = \frac{1}{2}\partial_\mu\chi\partial^\mu\chi +
460: \frac{1}{2}\partial_\mu\phi_a\partial^\mu\phi_a + \frac{R}{16\pi G} -
461: V(\phi,\chi)\,
462: \end{equation}
463: with $R$ the Ricci scalar. For hybrid models, we consider a generic
464: symmetry breaking `Higgs' field with $N_c$ real components. Thus, we
465: can take $\Phi^{\dagger}\Phi = {1\over2}\sum_a\phi_a^2 \equiv
466: |\phi|^2/2$ in (\ref{higgsPotential}), with $a$ running for the number
467: of Higgs' components, \textit{e.g.} $N_c = 1$ for a real scalar Higgs,
468: $N_c = 2$ for a complex scalar Higgs or $N_c = 4$ for a $SU(2)$ Higgs,
469: etc. The effective potential~(\ref{higgsPotential}) then becomes
470: \begin{equation}\label{higgsPotentialII}
471: \indent V(\phi,\chi) = \frac{\lambda}{4}\left(|\phi|^2 - 
472: v^2\right)^2 + g^2\chi^2|\phi|^2
473: + \frac{1}{2}\mu^2\chi^2\,.
474: \end{equation}
475: For chaotic scenarios, we 
476: consider a massless scalar field interacting with the 
477: inflaton via
478: \begin{equation}
479: \indent V(\chi,\phi) = \frac{1}{2}g^2\chi^2\phi^2 + V(\chi)\,,
480: \end{equation}
481: with $V(\chi)$ the inflaton's potential. Concerning the 
482: simulations we show in this paper, we concentrate in the 
483: $N_c = 4$ case for the hybrid model and consider a potential 
484: $V(\chi) = {\lambda\over4}\chi^4$ for the chaotic scenario.
485: 
486: The classical equations of motion of the
487: inflaton and the other Bose fields are
488: \begin{eqnarray}
489: \label{inflatonEq}
490: &&\ddot\chi + 3H\dot\chi - \frac{1}{a^2}\nabla^2\chi + \frac{\partial
491: V}{\partial\chi} = 0 \\
492: \label{scalarEq}
493: &&\ddot\phi_a + 3H\dot\phi_a - \frac{1}{a^2}\nabla^2\phi_a +
494: \frac{\partial V}{\partial\phi_a} = 0
495: \end{eqnarray}
496: with $H = \dot a/a$. 
497: 
498: Gravitational Waves are represented here by a transverse-traceless 
499: (TT) gauge-invariant metric perturbation,
500: $h_{ij}$, on top of the flat FRW space
501: \begin{equation}
502: \indent ds^2 = -dt^2 + a^2(t)\left(\delta_{ij} + 
503: h_{ij}\right)dx^idx^j\,,
504: \end{equation}
505: with $a(t)$ the scale factor and the tensor perturbations verifying 
506: $\partial_ih_{ij} = h_{ii} = 0$. In the following, we will raise 
507: or low indices of the metric perturbations with the delta Kronecker 
508: $\delta_{ij}$, so $h_{ij} = h^{i}_{j} = h^{ij}$ and so on.
509: The Einstein field equations 
510: can be splitted into the
511: background $G_{\mu\nu}^{(0)} = 8\pi G\,T_{\mu\nu}^{(0)}$ and the
512: perturbed $\delta G_{\mu\nu} = 8\pi G\,\delta {\rm T}_{\mu\nu}$
513: equations. The background equations describe the evolution of the flat
514: FRW universe through
515: \begin{eqnarray}
516: \label{hubbleDotEq}
517: -\frac{\dot H}{4\pi G} &=&  \dot\chi^2 +
518: \frac{1}{3a^2}(\nabla\chi)^2 + \dot\phi_a^2 +
519: \frac{1}{3a^2}(\nabla\phi_a)^2\\
520: \label{hubbleEq}
521: \frac{3H^2}{4\pi G} &=& \dot\chi^2 +
522: \frac{1}{a^2}(\nabla\chi)^2 + \dot\phi_a^2 +
523: \frac{1}{a^2}(\nabla\phi_a)^2 + 2V(\chi,\phi)\, \nonumber \\
524: \end{eqnarray}
525: where any term in the r.h.s. of~(\ref{hubbleDotEq}) and~(\ref{hubbleEq}), 
526: should be understood as spatially averaged.
527: 
528: On the other hand, the perturbed Einstein equations describe the evolution 
529: of the tensor perturbations~\cite{Mukhanov} as
530: \begin{equation}\label{GWeq}
531: \indent \ddot h_{ij} + 3H\dot h_{ij} - \frac{1}{a^2}\nabla^2h_{ij} =
532: 16\pi G\,\Pi_{ij}\,,
533: \end{equation}
534: with $\partial_i\Pi_{ij} = \Pi_{ii} = 0$. 
535: The source of the GW, $\Pi_{ij}$, 
536: contributed by both the inflaton and the other scalar fields, 
537: will be just the transverse-traceless part of the (spatial-spatial) 
538: components of the total anisotropic stress-tensor
539: \begin{equation}\label{ast}
540: \indent {\rm T}_{\mu\nu} = \frac{1}{a^2}\left[ \partial_\mu\chi
541: \partial_\nu\chi + \partial_\mu\phi_a\partial_\nu\phi_a + 
542: g_{\mu\nu}(\mathcal{L} - \left\langle p \right\rangle)\right]\,,
543: %\nonumber \\ 
544: \end{equation}
545: where $\mathcal{L}(\chi,\phi_a)$ is the lagrangian~(\ref{lagrangian})
546: and $\left\langle p \right\rangle$ is the background homogeneous
547: pressure. As we will explain in the next subsection, when extracting
548: the TT part of~(\ref{ast}), the term proportional to $g_{\mu\nu}$ in
549: the r.h.s of~(\ref{ast}), will be dropped out from the GW equations
550: of motion. Thus, the effective source of the GW will be just
551: given by the TT part of the gradient terms
552: $\partial_\mu\chi\partial_\nu\chi +
553: \partial_\mu\phi_a\partial_\nu\phi_a$.
554: 
555: In summary, Eqs.~(\ref{inflatonEq})-(\ref{scalarEq}), together with
556: Eqs.~(\ref{hubbleDotEq})-(\ref{hubbleEq}), describe the coupled
557: dynamics of reheating in any inflationary scenario, while
558: Eq.~(\ref{GWeq}) describe the production of GW in each of those
559: scenarios. In this paper we use lattice simulations to study the
560: generation of GW during reheating after inflation. Specific details on
561: this are given in section~IV, but let us just mention here that our
562: approach is to solve the evolution of the gravitational waves
563: simultaneously with the dynamics of the scalar fields, in a
564: discretized lattice with periodic boundary conditions. We assume
565: initial quantum fluctuations for all fields and only a zero mode for
566: the inflaton.  Moreover, we also included the GW backreaction on the
567: scalar fields' evolution via the gradient terms,
568: $h^{ij}\nabla_i\chi\nabla_j\chi + h^{ij}\nabla_i\phi_a\nabla_j\phi_a$
569: and confirmed that, for all practical purposes, these are negligible
570: throughout GW production.
571: 
572: \subsection{The Transverse-Traceless Gauge}
573: 
574: A generic (spatial-spatial) metric perturbation $\delta h_{ij}$ has six 
575: independent degrees of freedom, whose contributions can be split into 
576: scalar, vector and tensor metric perturbations~\cite{Mukhanov}
577: \begin{equation}
578: \indent
579: \delta h_{ij} = \psi\,\delta_{ij} + E_{,ij} + F_{(i,j)} + h_{ij}\,.
580: \end{equation}
581: with $\partial_iF_i = 0$ and $\partial_ih_{ij} = h_{ii} = 0$.  By
582: choosing a transverse-traceless stress-tensor source $\Pi_{ij}$, we
583: can eliminate all the degrees of freedom (d.o.f.) but the pure TT
584: part, $h_{ij}$, which represent the only physical d.o.f which
585: propagate and carry energy out of the source (GW). If we had chosen
586: only a traceless but non-transverse stress source, we could have
587: eliminated the scalar d.o.f.  $\psi$ and absorbed $E$ into the scalar
588: field perturbation, but we would still be left with a vector field
589: $F_i$ also sourced by the (traceless but non-transverse) anisotropic
590: stress tensor, thus giving rise to a vorticity divergence-less field
591: $V_i$. However, since the initial conditions are those of a scalar
592: Gaussian random field (see section IV), even in that case of a
593: non-transverse but traceless stress source, the mean vorticity of the
594: subsequent matter distribution, averaged over a sufficiently large
595: volume, should be zero (although locally we do have vortices of the
596: Higgs field, see Refs.~\cite{CEWB,magnetic}), since vortices with
597: opposite chirality cancel eachother. This means that in such a case,
598: although $\partial^i\Pi_{ij} \neq 0$, and thus $\partial^i\delta
599: h_{ij} \neq 0$, we could still recover the TT component when averaging
600: over the whole realization.
601: 
602: For practical purposes, we will consider from the beggining the TT
603: part of the anisotropic stress-tensor, ensuring this way that we only
604: source the physical d.o.f. that represent GW. The equations of motion
605: of the TT metric perturbations are then given by Eq.~(\ref{GWeq}).
606: Note that for a non-transverse source the equations would have been
607: much more complicated, so the advantage of using the TT part from the
608: beginning is clear. The disadvantage arises because obtaining the TT
609: part of a tensor in configuration space is very demanding in
610: computational time. However, as we explain next, we will use a method by
611: which we can circunvent this issue. 
612: 
613: Let us switch to Fourier space. Using the convention
614: \begin{equation}
615: \indent f(\mathbf{k}) = \frac{1}{(2\pi)^{3/2}}\int d^3\mathbf{x}\,
616: e^{+i\mathbf{kx}}f(\mathbf{x})\,,
617: \end{equation} 
618: the GW equations~(\ref{GWeq}) in Fourier space read
619: \begin{equation}\label{GWeqFourier}
620: \indent \ddot h_{ij}(t,\mathbf{k}) + 3H\dot h_{ij}(t,\mathbf{k}) + 
621: \frac{k^2}{a^2}h_{ij}(t,\mathbf{k}) = 16\pi G\,\Pi_{ij}(t,\mathbf{k})\,,
622: \end{equation}
623: where $k = |\mathbf{k}|$. Assuming no GW at the beginnig of reheating
624: (i.e.  the end of inflation $t_e$), the initial conditions are
625: $h_{ij}(t_e) = \dot h_{ij}(t_e) = 0$, so the solution to
626: Eq.~(\ref{GWeqFourier}) for $t > t_e$ will be just given by a causal
627: convolution with an appropriate green function $G(t,t')$,
628: \begin{eqnarray}\label{GWsol}
629: \indent h_{ij}(t,\mathbf{k}) = 16\pi G \int_{t_e}^{t}dt'\,G(t,t')
630: \Pi_{ij}(t',\mathbf{k})
631: \end{eqnarray} 
632: Therefore, all we need to know for computing the GW is the TT part of
633: the stress-tensor, $\Pi_{ij}$, and the Green function $G(t',t)$.
634: However, as we will demonstrate shortly, we have used a numerical
635: method by which we don't even need to know the actual form of
636: $G(t',t)$.  To see this, let us extract the TT part of the total
637: stress-tensor. Given the symmetric anisotropic stress-tensor
638: T$_{\mu\nu}$~(\ref{ast}), we can easily obtain the TT part of its
639: spatial components in momentum space, $\Pi_{ij}(\mathbf{k})$. Using
640: the spatial projection operators $P_{ij} = \delta_{ij} - \hat k_i \hat
641: k_j$, with $\hat k_i = k_i/k$, then~\cite{Carroll}
642: \begin{eqnarray}
643: \label{TTgauge}
644: \indent \Pi_{ij}(\mathbf{k}) = \Lambda_{ij,lm}(\mathbf{\hat k})
645: {\rm T}_{lm}(\mathbf{k})\,,
646: \end{eqnarray} 
647: where
648: \begin{eqnarray}\label{projector}
649: \indent \Lambda_{ij,lm}(\mathbf{\hat k}) \equiv \Big(P_{il}(\mathbf{\hat k})
650: P_{jm}(\mathbf{\hat k}) - \half P_{ij}(\mathbf{\hat k})
651: P_{lm}(\mathbf{\hat k})\Big)\,.
652: \end{eqnarray}
653: Thus, one can easily see that, at any time $t$, $k_i\Pi_{ij}(\mathbf{\hat
654: k},t) = \Pi_{i}^{i}(\mathbf{\hat k},t) = 0$, as required, thanks to
655: the identities $P_{ij}\hat k_j = 0$ and $P_{ij}P_{jm} = P_{im}$.
656: 
657: Note that the TT tensor, $\Pi_{ij}$, is just a linear combination of
658: the components of non-traceless nor-transverse tensor
659: T$_{ij}$~(\ref{ast}), while the solution~(\ref{GWsol}) is just linear
660: in $\Pi_{ij}$. Therefore, we can write the TT tensor perturbations 
661: (i.e. the gravitational waves) as
662: \begin{eqnarray}
663: \label{TTsol}
664: \indent h_{ij}(t,\mathbf{k}) = \Lambda_{ij,lm}(\mathbf{\hat k}) 
665:  u_{ij}(t,\mathbf{k})\,,
666: \end{eqnarray} 
667: with $u_{ij}(t,\mathbf{k})$ the Fourier transform of the solution 
668: of the following equation
669: \begin{eqnarray}\label{GWfakeEq}
670: \indent \ddot u_{ij} + 3H\dot u_{ij} - 
671: \frac{1}{a^2}\nabla^2 u_{ij} = 16\pi G\,{\rm T}_{ij}
672: \end{eqnarray}
673: This Eq.~(\ref{GWfakeEq}) is nothing but Eq.~(\ref{GWeq}), sourced
674: with the complete T$_{ij}$~(\ref{ast}), instead of with its TT part,
675: $\Pi_{ij}$.  Of course, Eq.~(\ref{GWfakeEq}) contains unphysical
676: (gauge) d.o.f.; however, in order to obtain the real physical TT
677: d.o.f., $h_{ij}$, we can evolve Eq.(\ref{GWfakeEq}) in configuration
678: space, Fourier transform its solution and apply the
679: projector~(\ref{projector}) as in ~(\ref{TTsol}). This way we can
680: obtain in momentum space, at any moment of the evolution, the physical
681: TT d.o.f. that represent GW, $h_{ij}$. Whenever needed, we can Fourier
682: transform back to configuration space and obtain the spatial
683: distribution of the gravitational waves.
684: 
685: Moreover, since the second term of the r.h.s of the total
686: stress-tensor T$_{ij}$ is proportional to $g_{ij} = \delta_{ij} +
687: h_{ij}$, see~(\ref{ast}), when aplying the TT
688: projector~(\ref{projector}), the part with the $\delta_{ij}$ just
689: drops out, simply because it is a pure trace, while the other part 
690: contributes with a term $-(\mathcal{L} - \left\langle p
691: \right\rangle)h_{ij}$ in the l.h.s of Eq.(\ref{GWeqFourier}). 
692: %Since
693: %$(\mathcal{L} - \left\langle p \right\rangle) = - (16\pi
694: %G/3)\left\langle\partial_i\chi\partial^i\chi + \partial_i\phi^a
695: %\partial^i\phi_a\right\rangle$,
696: However, $(\mathcal{L} - \left\langle p \right\rangle)$ is of the same 
697: order as the metric perturbation $\sim {\cal O}(h)$, 
698: so this extra term is second order in the gravitational coupling 
699: and it can be neglected in the GW Eqs.~(\ref{GWeqFourier}).
700: This way, the effective source in Eq.~(\ref{GWfakeEq}) is just the 
701: gradient terms of both the inflaton and the other scalar fields,
702: \begin{eqnarray}\label{GWfakeSource}
703: \indent {\rm T}_{ij} = \frac{1}{a^2}\left(\nabla_i\chi\nabla_j\chi + 
704: \nabla_i\phi_a\nabla_j\phi_a\right)
705: \end{eqnarray} 
706: Therefore, the effective source of the physical GW, will be just the
707: TT part of~(\ref{GWfakeSource}), as we had already mentioned before.
708: 
709: Alternatively, one could evolve the equation of the TT tensor
710: perturbation in configuration space, Eq.~(\ref{GWeq}), with the source
711: given by
712: \begin{equation}\label{GWtrueSource}
713: \Pi_{ij}({\bf x},t) = {1\over(2\pi)^{3/2}}\int d^3{\bf k}\,
714: e^{-i{\bf kx}}\Lambda_{ij,lm}({\bf \hat k}){\rm T}_{lm}({\bf k},t)\,, 
715: \end{equation}
716: such that $\partial_i\Pi_{ij}({\bf x},t) = \Pi_{ii}({\bf x},t) = 0$ at
717: any time.  So, at each time step of the evolution of the fields, one
718: would have first to compute (the gradient part of)
719: T$_{lm}$~(\ref{GWfakeSource}) in configuration space, then Fourier
720: transform it to momentum space, substitute in Eq.~(\ref{GWtrueSource})
721: and perform the integral. However, proceeding as we suggested above,
722: there is no need to perform the integral,
723: %\footnote{Of course, if we want to obtain the spatial distribution
724: %of the tensor perturbations, we have to perform an inverse Fourier
725: %transform from $h_{ij}({\bf k},t)$ to $h_{ij}({\bf x},t)$. However, we
726: %can do this only at those times when we are interested on taking a
727: %snapshot of the GW spatial distribution, but not at every time step of
728: %the evolution of the fields.}, 
729: nor Fourier transform the fields at each time step, but rather only at
730: those times at which we want to measure the GW spectrum. The viability
731: of our method relies in the following observation. To compute the GW
732: we could, first of all, project the TT part of the
733: source~(\ref{GWtrueSource}), and second, solve Eq.~(\ref{GWeq}).
734: However, we achieve the same result if we commute these two operations
735: such that, first, we solve Eq.~(\ref{GWfakeEq}), and second, we apply
736: the TT projector to the solution~(\ref{TTsol}). We have found this
737: \textit{commuting procedure} very useful, since we are able to extract
738: the spectra or the spatial distribution of the GW at any desired time,
739: saving a great amount of computing time. Most importantly, with this
740: procedure we can take into account backreaction simultaneously with
741: the fields evolution.
742: 
743: In summary, for solving the dynamics of reheating of a particular
744: inflationary model, we evolve Eqs.~(\ref{inflatonEq})-(\ref{scalarEq})
745: in the lattice, together with
746: Eqs.~(\ref{hubbleDotEq})-(\ref{hubbleEq}), while for the GWs 
747: %, instead of solving Eq.~(\ref{GWeq}), 
748: we solve Eq.~(\ref{GWfakeEq}). Then, only
749: when required, we Fourier transform the solution of
750: Eq.~(\ref{GWfakeEq}) and then apply~(\ref{TTsol}) in order to recover
751: the physical transverse-traceless d.o.f representing the GW. From
752: there, one can easily build the GW spectra or take a snapshot of
753: spatial distribution of the gravitational waves.
754: 
755: \subsection{The energy density in GW}
756: 
757: The energy-momentum tensor of the GW is given by~\cite{Carroll}
758: \begin{equation}\label{tmunu}
759: \indent t_{\mu\nu} = {1\over 32\pi G}\,\left\langle\partial_\mu
760: h_{ij}\, \partial_\nu h^{ij}\right\rangle_{\rm V}\,,
761: \end{equation}
762: where $h_{ij}$ are the TT tensor perturbations solution of
763: Eq.(\ref{GWeq}). The expectation value
764: $\left\langle...\right\rangle_{\rm V}$ is taken over a region of
765: sufficiently large volume $V=L^3$ to encompass enough physical
766: curvature to have a gauge-invariant measure of the GW energy-momentum
767: tensor.
768: 
769: The GW energy density will be just 
770: $\rho_{_{\rm GW}} = t_{00}$, so
771: \begin{eqnarray}\label{rho}
772: \indent \rho_{GW} &=& \frac{1}{32\pi G}\frac{1}{L^3}\int d^3\mathbf{x}\,
773: \dot h_{ij}(t,\mathbf{x})\dot h_{ij}(t,\mathbf{x}) \nonumber \\ 
774: &=&  \frac{1}{32\pi G}\frac{1}{L^3}\int d^3\mathbf{k}\,
775: \dot h_{ij}(t,\mathbf{k})\dot h_{ij}^*(t,\mathbf{k})\,,
776: \end{eqnarray} 
777: where in the last step we Fourier transformed each 
778: $\dot h_{ij}$ and used the integral definition of the Dirac delta 
779: $(2\pi)^3\delta^{(3)}(\mathbf{k}) = \int d^3\mathbf{x}\ e^{-i\mathbf{kx}}$.
780: 
781: We can always write the scalar product in (\ref{rho}) in terms of the 
782: (Fourier transformed) solution $u_{lm}$ of the Eq.(\ref{GWfakeEq}), 
783: by just using the
784: spatial projectors (\ref{projector})
785: \begin{equation}
786: \indent \dot h_{ij}\dot h_{ij} = \Lambda_{ij,lm}\Lambda_{ij,rs} 
787: \dot u_{lm} \dot u_{rs} = 
788: \Lambda_{lm,rs} \dot u_{lm} \dot u_{rs}\,,
789: \end{equation}
790: where we have used the fact that $\Lambda_{ij,lm}\Lambda_{lm,rs} =
791: \Lambda_{ij,rs}$. This way, we can express the GW energy density as
792: \begin{eqnarray}\label{rhoTotal}
793: &&\rho_{GW} = \frac{1}{32\pi G L^3}\times \nonumber \\ 
794: &&\hspace{1.2cm} \int k^2dk \int d\Omega\,\Lambda_{ij,lm}(\mathbf{\hat k})
795: \dot  u_{ij}(t,\mathbf{k})\dot  u_{lm}^*(t,\mathbf{k}).
796: \end{eqnarray}
797: From here, we can also compute the power spectrum per logarithmic frequency
798: interval in GW, normalized to the critical density $\rho_c$, as
799: \begin{equation}
800: \indent \Omega_{_{\rm GW}}=\int {df\over f}\,\Omega_{_{\rm GW}}(f)\,, 
801: \end{equation}
802: where
803: \begin{eqnarray}\label{OmegaFraction}
804: \Omega_{gw}(k) \equiv \frac{1}{\rho_c}\frac{d\rho_{gw}}{d\,{\rm log}k}\hspace{5.2cm} \nonumber \\
805: = \frac{k^3}{32\pi GL^3\rho_c} \int d\Omega\,\Lambda_{ij,lm}(\mathbf{\hat k})
806: \dot  u_{ij}(t,\mathbf{k})\dot  u_{lm}^*(t,\mathbf{k})\hspace{.4cm} 
807: \end{eqnarray}
808: We have checked explicitely in the simulations that the argument 
809: of the angular integral of~(\ref{OmegaFraction}) is independent 
810: of the directions in {\bf k}-space. Thus, whenever we plot the GW 
811: spectrum of any model, we will be showing the amplitude of the 
812: spectrum (per each mode $k$) as obtained after avaraging over all 
813: the the directions in momentum space,
814: \begin{eqnarray}
815: \indent\Omega_{gw}(k) = \frac{k^3}{8GL^3\rho_c} \left\langle \Lambda_{ij,lm}(\mathbf{\hat k})
816: \dot  u_{ij}(t,\mathbf{k})\dot  u_{lm}^*(t,\mathbf{k})\right\rangle_{4\pi} 
817: \end{eqnarray} 
818: where $\left\langle f \right\rangle_{4\pi} \equiv {1\over 4\pi}\int f{\rm d}\Omega$. 
819: 
820: Finally, we must address the fact that the frequency range, for a GWB
821: produced in the early universe, will be redshifted today. We should
822: calculate the characteristic physical wavenumber of the present GW
823: spectrum, which is redshifted from any time $t$ during GW
824: production. This is a key point, since a relatively long period of
825: turbulence will develop after preheating, which could change the
826: amplitude of the GWB and shift the frequency at which the spectra
827: peaks. So let us distinguish four characteristic times: the end of
828: inflation, $t_e$; the time $t_*$ when GW production stops;
829: the time $t_{r}$ when the universe finally reheats and enters into the
830: radiation era; and today, $t_0$.\footnote{Note, however, that after
831: thermalization there is still a small production of GW from the
832: thermal plasma, but this can be ignored for all practical purposes.}
833: Thus, today's frequency $f_0$ is related to the physical wavenumber
834: $k_t$ at any time $t$ of GW production, via $f_0 =
835: (a_t/a_0)(k_t/2\pi)$, with $a_0$ and $a_t$, the scale factor today and
836: at the time $t$, respectively. Thermal equilibrium was established at
837: some temperature $T_r$, at time $t_r \geq t$. The Hubble rate at that
838: time was $M_P^2H_r^2 = (8\pi/3)\rho_r$, with $\rho_r =
839: g_r\pi^2T_r^4/30$ the relativistic energy density and $g_r$ the
840: effective number of relativistic degrees of freedom at temperature
841: $T_r$. Since then, the scale factor has increased as $a_r/a_0 =
842: (g_{0,s}/g_{r,s})^{1/3}(T_0/T_{r})$, with $g_{i,s}$ the effective
843: entropic degrees of freedom at time $t_i$, and $T_0$
844: today's CMB temperature. Putting all together,
845: \begin{eqnarray}
846: \label{redshift}
847: \indent f_0 = \left(\frac{8\pi^3g_{r}}{90}\right)^{1\over4}
848: \left(\frac{g_{0,s}}{g_{{r},s}}\right)^{1\over3}\frac{T_0}
849: {\sqrt{H_{r}M_p}}\left(\frac{a_e}{a_r}\right)\frac{k}{2\pi}\,,
850: \end{eqnarray}
851: where we have used the fact that the physical wave number $k_t$ at 
852: any time $t$ during GW production, is related to the comoving wavenumber $k$
853: through $k_t = (a_e/a_t)k$ with the normalization $a_e \equiv 1$.
854: 
855: From now on, we will be concerned with hybrid inflation, leaving
856: chaotic inflation for section VI. Within the hybrid scenario, we will
857: analyse the dependence of the shape and amplitude of the produced GWB
858: on the scale of hybrid inflation, and more specifically on the
859: \textit{v.e.v.} of the Higgs field triggering the end of inflation.
860: The initial conditions are carefully treated following the
861: prescription adopted in paper I, as explained in section IV. Given the
862: natural frequency at hand in hybrid models, $m = \sqrt{\lambda}v$,
863: whose inverse $m^{-1}$ sets the characteristic time scale during the
864: first stages of reheating, it happens that as long as $v\,\ll\,M_p$,
865: the Hubble rate $H\,\sim\,\sqrt{\lambda}(v^2/M_p)$ is much smaller
866: than such a frequency, $H\,\ll\,m$. Indeed, all the initial vacuum
867: energy $\rho_0$ gets typically converted into radiation in less than a
868: Hubble time, in just a few $m^{-1}$ time steps. Therefore, we should
869: be able to ignore the dilution due to the expansion of the universe
870: during the production of GW, at least during the first stages of
871: reheating. However, as we will see, the turbulent behaviour developed
872: after those first stages, could last for much longer than an e-fold,
873: in which case we will have to take into account the expansion of the
874: universe. Our approach will be first to ignore the expansion of the
875: Universe and later see how we can account for corrections if
876: needed. Thus, we set the scale factor $a = 1$ and the Hubble rate $H =
877: 0$ and $\dot H = 0$. As we will see later, our approach of neglecting 
878: the expansion for the time of GW production, will be completely justified
879: \textit{a~posteriori}.
880: 
881: The coupled evolution equations that
882: we have to solve numerically in a lattice for the hybrid model are
883: \begin{eqnarray}\label{GWeqHybrid}
884: &&\ddot u_{ij} - \nabla^2 u_{ij} = 16\pi G\,{\rm T}_{ij} \\
885: &&\ddot\chi - \nabla^2\chi + \left(g^2|\phi|^2 +
886: \mu^2\right)\chi = 0 \\ &&\ddot\phi_a - \nabla^2\phi_a +
887: \left(g^2\chi^2 + \lambda|\phi|^2 - m^2\right)\phi_a = 0
888: \end{eqnarray}
889: with T$_{ij}$ given by Eq.(\ref{GWfakeSource}) with the scale factor
890: $a = 1$. We have explicitly checked in our computer simulations that
891: the backreaction of the gravity waves into the dynamics of both the
892: inflaton and the Higgs fields is negligible and can be safely ignored.
893: We thus omit the backreaction terms in the above equations.
894: 
895: 
896: %%%%%%%%%%%%%%%%%%%%%%%%%
897: \begin{figure}[t]
898: %\vspace{5mm}
899: \begin{center}
900: %\includegraphics[width=4.cm,height=7.cm,angle=-90]{meanFields.eps}
901: \includegraphics[width=5.5cm,height=8.5cm,angle=-90]{meanFields.eps}
902: \end{center}
903: \vspace*{-5mm}
904: \caption{Time evolution of the mean field values of the Higgs and the
905: Inflaton, the former normalized to its \textit{v.e.v.}, the latter
906: normalized to its critical value $\chi_{0} = \rm{m/g}$. This is just a
907: specific realization with $N = 128$, $p_{\rm min} = 0.1m$, $a =
908: 0.48m^{-1}$, $v = 10^{-3}M_p$ and $g^2 = 2\lambda = 0.25$.}
909: \label{fig1}
910: \vspace*{-3mm}
911: \end{figure}
912: %%%%%%%%%%%%%%%%%%%%%%%%%
913: 
914: We evaluate during the evolution of the system the mean field values,
915: as well as the different energy components. As shown in
916: Fig.~\ref{fig1}, the Higgs field grows towards the true vacuum and
917: the inflaton moves towards the minimum of its potential and oscillates
918: around it. We have checked that the sum of the averaged gradient,
919: kinetic and potential energies (contributed by both the inflaton and
920: the Higgs), remains constant during reheating, as expected, since the
921: expansion of the universe is irrelevant in this model. We have also
922: checked that the time evolution of the different energy components is
923: the same for different lattices, i.e. changing the number of points
924: $N$ of the lattice, of the minimum momentum $p_{\rm min} = 2\pi/L$ or
925: of the lattice spacing $a = L/N$, with $L$ the lattice size, as long
926: as the product $ma < 0.5$; for a detailed discussion of lattice scales
927: see paper I.  Looking at the time evolution of the Higgs'
928: \textit{v.e.v.}  in  Fig.~\ref{fig1}, three stages can be
929: distinguished. First, an exponential growth of the \textit{v.e.v.}
930: towards the true vacuum. This is driven by the tachyonic instability
931: of the long-wave modes of the Higgs field, that makes the spatial
932: distribution of this field to form lumps and bubble-like
933: structures~\cite{tachyonic,symmbreak}. Second, the Higgs field
934: oscillates around the true vacuum, as the Higgs' bubbles collide and
935: scatter off eachother. Third, a period of turbulence is reached,
936: during which the inflaton oscillates around its minimum and the Higgs
937: sits in the true vacuum. For a detailed description of the dynamics of
938: these fields see Ref.~\cite{symmbreak}. Here we will be only concerned
939: with the details of the gravitational wave production.
940: 
941: %%%%%%%%%%%%%%%%%%%%%%%%%
942: \begin{figure}[t]
943: %\vspace{5mm}
944: \begin{center}
945: \includegraphics[width=5.5cm,height=8.5cm,angle=-90]{energygw.ps}
946: \end{center}
947: \vspace*{-5mm}
948: \caption{The time evolution of the different types of energy (kinetic,
949: gradient, potential, anisotropic components and gravitational waves
950: for different lattices), normalized to the initial vacuum energy,
951: after hybrid inflation, for a model with $v=10^{-3}\,M_P$. One can
952: clearly distinguish here three stages: tachyonic growth, bubble
953: collisions and turbulence. }
954: \label{fig3}
955: \vspace*{-3mm}
956: \end{figure}
957: %%%%%%%%%%%%%%%%%%%%%%%%%
958: 
959: The initial energy density at the end of hybrid inflation is given by
960: $\rho_0 = m^2v^2/4$, with $m^2 = \lambda v^2$, so the fractional
961: energy density in gravitational waves is
962: \begin{equation}
963: \indent {\rho_{_{\rm GW}}\over\rho_0} = {4t_{00}\over v^2\,m^2} =
964: \frac{1}{8\pi\,G v^2 m^2}\left\langle\dot h_{ij}\dot
965: h^{ij}\right\rangle_{\rm V}\,,
966: \end{equation}
967: where $\left\langle\dot h_{ij}\dot h^{ij}\right\rangle_V$, defined as a 
968: volume average like 
969: ${1\over {\rm V}}\int d^3{\bf x}\dot h_{ij}\dot h^{ij}$, is extracted from 
970: the simulations as
971: \begin{eqnarray}
972: &&\left\langle\dot h_{ij}\dot
973: h^{ij}\right\rangle_{\rm V} = \nonumber \\ && = 
974: \dfrac{4\pi}{V}\int d{\rm log}k\,k^3\left\langle \Lambda_{ij,lm}(\mathbf{\hat k})
975: \dot  u_{ij}(t,\mathbf{k})\dot  u_{lm}^*(t,\mathbf{k})
976: \right\rangle_{4\pi} 
977: \end{eqnarray}
978: where $ u_{ij}(t,\mathbf{k})$ is the Fourier transform of the solution 
979: of Eq.~(\ref{GWeqHybrid}). 
980: Then, we can compute the corresponding
981: density parameter today (with $\Omega_{\rm
982: rad}\,h^2\simeq3.5\times10^{-5}$)
983: \begin{eqnarray}
984: && \Omega_{_{\rm GW}}\,h^2 = {\Omega_{\rm rad}\,h^2\over 2
985: G\,v^2\,m^2\,V}\,\times \nonumber \\ &&\hspace{.7cm} 
986: \int d{\rm log}k\, k^3\left\langle \Lambda_{ij,lm}(\mathbf{\hat k})
987: \dot  u_{ij}(t,\mathbf{k})\dot  u_{lm}^*(t,\mathbf{k})
988: \right\rangle_{4\pi} %\nonumber \\
989: \end{eqnarray}
990: which has assumed that all the
991: vacuum energy $\rho_0$ gets converted into radiation, an approximation
992: which is always valid in generic hybrid inflation models with $v\ll
993: M_P$, and thus $H\ll m=\sqrt\lambda\,v$.
994: 
995: We have shown in Fig.~\ref{fig3} the evolution in time of the fraction
996: of energy density in GW. The first (tachyonic) stage is clearly
997: visible, with a (logarithmic) slope twice that of the anisotropic
998: tensor $\Pi_{ij}$. Then there is a small plateau corresponding to the
999: production of GW from bubble collisions; and finally there is the
1000: slow growth due to turbulence. In the next section we will describe
1001: in detail the most significant features appearing at each stage.
1002: 
1003: Note that in the case that $H\ll m$, the maximal production of GW
1004: occurs in less than a Hubble time, soon after symmetry breaking, while
1005: turbulence lasts several decades in time units of $m^{-1}$. Therefore,
1006: we can safely ignore the dilution due to the Hubble expansion, up to
1007: times much greater than those of the tachyonic instability. Eventually
1008: the universe reheats and the energy in gravitational waves redshifts
1009: like radiation thereafter.
1010: 
1011: %%%%%%%%%%%%%%%%%%%%%%%%%
1012: \begin{figure}[t]
1013: \vspace{-4mm}
1014: \begin{center}
1015: \includegraphics[width=5.5cm,height=8.7cm,angle=-90]{spectraGW_TT_t100.eps}
1016: \end{center}
1017: \vspace*{-5mm}
1018: \caption{We show here the comparison between the power spectrum of
1019: gravitational waves obtained with increasing lattice resolution, to
1020: prove the robustness of our method. The different realizations are
1021: characterized by the %number of lattice points (N), 
1022: the minimum 
1023: lattice momentum (p$_{\rm min}$) and the lattice spacing (ma). The
1024: growth is shown in steps of $m\Delta t = 1$ up to $mt = 30$, and 
1025: % for the lower
1026: then in and $m\Delta t = 5$ steps up to $mt = 60$.}
1027: \label{fig4}
1028: \vspace*{-3mm}
1029: \end{figure}
1030: %%%%%%%%%%%%%%%%%%%%%%%%%
1031: 
1032: To compute the power spectrum per logarithmic frequency
1033: interval in GW, $\Omega_{gw}(f)$, we just have to use~(\ref{OmegaFraction}).
1034: %\begin{equation}
1035: %\indent \Omega_{_{\rm GW}}(k)={k^3\over2\pi^2}\,{\rho_{_{\rm GW}}(k)
1036: %\over\rho_c}\,, 
1037: %\end{equation}
1038: Moreover, since gravitational waves below Planck scale remain decoupled 
1039: from the plasma immediately after
1040: production, we can evaluate the power spectrum today from that
1041: obtained at reheating by converting the wavenumber $k$ into
1042: frequency $f$. Simply using Eq.~(\ref{redshift}), with 
1043: $g_{{r},s}/g_{0,s}\sim 100$, $g_{{r},s}\,\sim\,g_{r}$ and 
1044: $a_e \sim a_*$, then
1045: \begin{equation}
1046: \indent f=6\times10^{10}\,{\rm Hz}\,{k\over\sqrt{H\,M_p}}
1047: =5\times10^{10}\,{\rm Hz}\,{k\over m}\,\lambda^{1/4}\,.
1048: \end{equation}
1049: We show in Fig.~\ref{fig4} the power spectrum of gravitational waves
1050: as a function of (comoving) wavenumber $k/m$. We have used different
1051: lattices in order to have lattice artifacts under control, specially
1052: at late times and high wavenumbers. We made sure by the choice of
1053: lattice size and spacing (i.e. $k_{\rm min}$ and $k_{\rm max}$) that
1054: all relevant scales fitted within the simulation. Note, however, that
1055: the lower bumps are lattice artifacts, due to the physical cutoff
1056: imposed at the initial condition, that rapidly disappear with time. We
1057: have also checked that the power spectrum of the scalar fields follows
1058: turbulent scaling after $mt \sim {\cal O}(100)$, and we can thus
1059: estimate the subsequent evolution of the energy density distributions
1060: beyond our simulations.
1061: 
1062: \section{Lattice simulations}
1063: 
1064: The problem of determining the time evolution of a quantum field
1065: theory is an outstandingly difficult problem. In some cases only a few
1066: degrees of freedom are relevant or else perturbative techniques are
1067: applicable. However, in our particular case, our interests are focused
1068: on processes which are necessarily non-linear and non-perturbative and
1069: involve many degrees of freedom. The presence of gravitational fields
1070: just contributes with more degrees of freedom, but does not complicate 
1071: matters significantly.
1072: 
1073: The lattice formulation allows a first principles approach to
1074: non-perturbative quantum field theory. The existing powerful lattice
1075: field theory numerical methods rest on the path integral formulation
1076: in Euclidean space and the existence of a probability measure in field
1077: space~\cite{Wilson}. However, the problem we are interested in is a
1078: dynamical process far from equilibium, and the corresponding Minkowski
1079: path integral formulation is neither mathematically well founded nor
1080: appropriate for numerical studies. There are a series of alternative
1081: non-perturbative methods which different research groups have used to
1082: obtain physical results in situations similar to ours. These include
1083: Hartree's approximations~\cite{Cooper} to go beyond perturbation
1084: theory or large N techniques~\cite{Vega,Baacke}. It is clear that it
1085: is desirable to look at this and similar problems with all available
1086: tools. In the present paper we will use an alternative approximation
1087: to deal with the problem: the classical approximation.  It consists of
1088: replacing the quantum evolution of the system by its classical
1089: evolution, for which there are feasible numerical methods
1090: available. The quantum nature of the problem remains in the stochastic
1091: character of the initial conditions. This approximation has been used
1092: with great success by several groups in the
1093: past~\cite{classical,tachyonic}. The advantage of the method is that
1094: it is fully non-linear and non-perturbative, allows the use of gauge
1095: fields~\cite{CEWB,magnetic} and gives access to the quantities we are
1096: interested in.
1097:  
1098: The validity of the approximation depends on the loss of quantum
1099: coherence in the evolution of the system. In previous papers we
1100: studied this problem both in the absence of and with gauge
1101: fields~\cite{symmbreak,CEWB,magnetic}. We started the evolution of the
1102: system at the critical time $t_c$, corresponding to the end of
1103: inflation $t_e$, at which the effective mass of the Higgs vanishes,
1104: putting all the modes in its (free field) ground states. If the
1105: couplings are small, since the quantum fluctuations of the value of
1106: the fields are not too large, the non-linear terms in the Hamiltonian
1107: of the system can be neglected. Then the quantum evolution is Gaussian
1108: and can be studied exactly. The Hamiltonian has nonetheless a
1109: time-dependence coming through the time-dependence of the inflaton
1110: homogeneous mode. This time dependence can always be taken to be
1111: linear for a sufficiently short time interval after the critical
1112: time. As a consequence, the dynamics of the eigenmodes during this
1113: initial phase differs significantly from mode to mode. Most of them
1114: have a characteristic harmonic oscillator behaviour with a frequency
1115: depending on the mode in question.  In the case of the Higgs field,
1116: the long-wave modes become tachyonic.  By looking at expectation
1117: values of products of these fields at different times, one realises
1118: that after a while these modes behave and evolve like classical modes
1119: of an exponentially growing size. The process is very fast and
1120: therefore the remaining harmonic modes can be considered to have
1121: remained in their initial ground state.
1122:  
1123: The fast growth in size of the Higgs field expectation value boosts
1124: the importance of non-linear terms and eventually drives the system
1125: into a state where the non-linear dynamics, including the
1126: back-reaction to the inflaton field, are crucial. For the whole
1127: approximation to be useful this must happen at a later time than the
1128: one in which the low-frequency Higgs modes begin evolving as classical
1129: fields. In paper I we showed this to be the case.  Actually, there is
1130: a time interval in which classical behaviour is already dominant while
1131: non-linearities are still small. We tested that our results, in the
1132: absence of gauge fields, were insensitive to the matching time,
1133: provided it lies within this window.
1134:  
1135: The whole idea can then be summarised as follows: the tachyonic
1136: quantum dynamics of the low momentum Higgs modes drive them into
1137: classical field behaviour and large occupation numbers before the
1138: non-linearities and back-reaction begin to play a role. It is the
1139: subsequent non-linear classical behaviour of the field that induces
1140: the growth of classical inflaton and gravitational field components
1141: also at low frequencies.  It is obvious that the quantum nature of the
1142: problem is still crucial if one studies the behaviour of high momentum
1143: modes which have low occupation numbers. 
1144:  
1145: In the present paper we apply the same idea in the presence of
1146: (gravitational wave) tensor fields. The initial quantum evolution of
1147: tensor fields is also relatively slow, since there are no tachyonic
1148: modes. Therefore, it is assumed not to affect substantially the
1149: initial conditions of the classical system. 
1150: 
1151: 
1152: \subsection{Initial conditions}
1153: 
1154: Our approach to the dynamics of the system is to assume that the
1155: leading effects under study can be well-described by the classical
1156: evolution of the system. The justification of this point, as explained
1157: in the previous section, rests upon the fast quantum evolution of the
1158: long wavelength components of the Higgs field during the initial
1159: stages after the critical point. All the other degrees of freedom will
1160: evolve slowly from their initial quantum vacuum state. For the Higgs
1161: field, the leading behaviour is the exponential growth of those modes
1162: having negative effective mass-squared. The quantum evolution of such
1163: modes drives the system into a quasi-classical regime. It is essential
1164: that this regime is reached before the non-linearities couple all
1165: degrees of freedom to each other and questions like back-reaction
1166: start to affect the results. It is then assumed that it is the
1167: essentially classical dynamics of that field what matters, and that
1168: all the long-wavelength components of the inflaton and the gauge
1169: fields produced by the interaction with the Higgs field behave also as
1170: classical fields. Of course, this can hardly be the case for shorter
1171: wavelengths which stay in a quantum state with low occupation numbers.
1172: However, as we can see in Fig.~\ref{fig4}, for the range of times
1173: studied in this paper, the effect of shorter wavelengths is relatively
1174: small, and the spectrum of modes remains always dominated by
1175: long-wavelengths. 
1176: 
1177: The full non-linear evolution of the system can then be studied using
1178: lattice techniques. Our approach is to discretize the classical
1179: equations of motion of all fields in both space and time. The
1180: time-like lattice spacing $a_t$ must be smaller than the spatial one
1181: $a_s$ for the stability of the discretized equations. In addition to
1182: the ultraviolet cut-off one must introduce an infrared cut-off by
1183: putting the system in a box with periodic boundary conditions.  We
1184: have studied $64^3$, $128^3$ and $256^3$ lattices.  Computer memory
1185: and CPU resources limit us from reaching much bigger lattices.
1186: Nonetheless, in the spirit of paper I, there are a number of
1187: checks one can perform to ensure that our results are physical and are
1188: not biased, within errors, by the approximations introduced, see 
1189: Fig.~\ref{fig4}. Our problem has several physical scales which control
1190: different time-regimes and observables. Thus, it is not always an easy
1191: matter to place these scales in the window defined by our ultraviolet
1192: and infrared cut-offs.  For example, in addition to the Higgs and
1193: inflaton mass there is a scale $M$ associated to the inflaton velocity
1194: which is particularly relevant in determining the bubble sizes and
1195: collisions.  Then, when we want to study a stage of the evolution in
1196: particular, we make the selection of the volume and cutoff most
1197: suitable.
1198: 
1199: Since in this paper we are more interested in understanding the
1200: phenomenon of GW production, rather than concentrating in a particular
1201: model, our attitude has been to modify the parameters of the model in
1202: order to sit in a region where our results are insensitive to the
1203: cut-offs.  This is no doubt a necessary first step to determine the
1204: requirements and viability of the study of any particular model. In
1205: particular, we have thouroughly studied a model with
1206: $g^2=2\lambda=1/4$, but we have checked that other values of the
1207: parameters do not change our results significantly.
1208: 
1209: The initial conditions of the fields follow the prescription from
1210: paper I. The Higgs modes $\phi_k$ are solutions of the coupled
1211: evolution equations, which can be rewritten as
1212: $\phi''_k+(k^2-\tau)\phi_k=0$, with $\tau=M(t-t_c)$ and
1213: $M=(2V)^{1/3}m$. The time-dependent Higgs mass follows from the
1214: initial inflaton field homogeneous component, $\chi_0(t_i) =
1215: \chi_c(1-Vm(t_i-t_c))$ and $\dot\chi_0(t_i)=-\chi_cVm$. The Higgs
1216: modes with $k/M > \sqrt{\tau_i}$ are set to zero, while the rest are
1217: determined by a Gaussian random field of zero mean distributed
1218: according to the Rayleigh distribution
1219: \begin{equation}
1220: \indent 
1221: P(|\phi_k|)d|\phi_k|d\theta_k = \exp\left(-{|\phi_k|^2\over\sigma_k^2}
1222: \right)\,{d|\phi_k|^2\over\sigma_k^2}\,{d\theta_k\over2\pi}\,,
1223: \end{equation}
1224: with a uniform random phase $\theta_k\in[0,2\pi]$ and dispersion given 
1225: by $\sigma_k^2 \equiv |f_k|^2 = P(k,\tau_i)/k^3$, where $P(k,\tau_i)$ is the power 
1226: spectrum of the initial Higgs quantum fluctuations in the background 
1227: of the homogeneous inflaton, computed in the linear approximation.
1228: In the region of low momentum modes it is well approximated by
1229: \begin{equation}\label{Papp}
1230: \indent 
1231: 2kP_{\rm app}(k,\tau_i) = k^3\left(1+A(\tau_i)\,k^2\,e^{-B(\tau_i)\,k^2}\right)\,,
1232: \end{equation}
1233: where $A(\tau_i)$ and $B(\tau_i)$ are parameters extracted from a fit
1234: of this form to the exact power spectrum given in paper~I. In the
1235: classical limit, the conjugate momentum $\dot\phi_k(\tau)$ is uniquely
1236: determined through $\dot\phi_k(\tau) = F(k,\tau)\phi_k(\tau)$, where
1237: $F(k,\tau) = {\rm Im}(if_k(\tau)\dot f_k^*(\tau))/|f_k(\tau)|^2$, see
1238: paper I. In the region of low momenta, $F(k,\tau_i)$ can be well
1239: aproximated by
1240: \begin{eqnarray}
1241: \indent F(k,\tau_i) = \frac{2kC(\tau_i)e^{-D(\tau_i)k^2}}{[1+A(\tau_i)e^{-B(\tau_i)k^2}]}.
1242: \end{eqnarray}
1243: where $A(\tau_i)$ and $B(\tau_i)$ are the previous coefficients for the amplitude of the field fluctuations, while $C(\tau_i)$ and $D(\tau_i)$ are new coefficients obtained fitting the exact expression of $F(k,\tau_i)$.
1244: 
1245: %%%%%%%%%%%%%%%%%%%%%%%%%
1246: \begin{figure}[t]%[t]
1247: %\vspace{5mm}
1248: \begin{center}
1249: \includegraphics[width=5.5cm,height=8.5cm,angle=-90]{HiggsSpectrum.ps}
1250: \end{center}
1251: \vspace*{-5mm}
1252: \caption{The tachyonic growth of the Higgs' spectrum, from $mt=5$ to
1253: $mt=10$. We compare simulations of different sizes ($p_{\rm min} = 
1254: 0.01 - 0.03$) and $N=256$, with the anaylitical expressions.} 
1255: \label{HigSpec}
1256: \vspace*{-3mm}
1257: \end{figure}
1258: %%%%%%%%%%%%%%%%%%%%%%%%%
1259: 
1260: The rest of the fields (the inflaton non-zero modes and the
1261: gravitational waves), are supposed to start from the vacuum, and
1262: therefore they are semiclassically set to zero initially in the
1263: simulations. Their coupling to the Higgs modes will drive their
1264: evolution, giving rise to a rapid (exponential) growth of the GW and
1265: inflaton modes.  Their subsequent non-linear evolution will be well
1266: described by the lattice simulations.
1267: 
1268: In the next subsections we will describe the different evolution 
1269: stages found in our simulations.
1270: 
1271: 
1272: \subsection{Tachyonic growth}
1273: 
1274: 
1275: In this subsection we will compare the analytical estimates with our
1276: numerical simulations for the initial tachyonic growth of the Higgs
1277: modes and the subsequent growth of gravitational waves. The first
1278: check is that the Higgs modes grow according to paper I.
1279: There we found that 
1280: \begin{equation}\label{kphi2}
1281: \indent 
1282: k|\phi_k(t)|^2 \simeq v^2\,A(\tau)\,e^{-B(\tau)k^2}\,, 
1283: \end{equation}
1284: with $A(\tau)$ and $B(\tau)$ are given in paper I,
1285: \begin{equation}\label{ABT}
1286: \indent 
1287: A(\tau) = {\pi^2(1/3)^{2/3}\over2\Gamma^2(1/3)}\,{\rm Bi}^2(\tau)\,,
1288: \hspace{5mm}
1289: B(\tau) = 2(\sqrt\tau - 1)\,,
1290: \end{equation}
1291: which are valid for $\tau>1$, and where ${\rm Bi}(z)$ is the Airy 
1292: function of the second kind. Indeed, we can
1293: see in Fig.~\ref{HigSpec} that the initial growth, from $mt=6$ to
1294: $mt=10$, follows precisely the analytical expression, once taken into
1295: account that in Eq.~(\ref{kphi2}) the wavenumber $k$ and time $\tau$ are
1296: given in units of $M=(2V)^{1/3}m$.
1297: 
1298: 
1299: 
1300: %%%%%%%%%%%%%%%%%%%%%%%%%
1301: \begin{figure}[t]%[t]
1302: %\vspace{5mm}
1303: \begin{center}
1304: \includegraphics[width=5.5cm,height=8.5cm,angle=-90]{P11Spectrum.ps}
1305: \end{center}
1306: \vspace*{-5mm}
1307: \caption{The Fourier transform of the anisotropic stress tensor.  We
1308: compare the numerical simulations of $\Pi_{11}(k,t)$ for $p_{\rm min}
1309: = 0.01$ with the analytical expressions (dashed lines) for $mt=5 -
1310: 10$, i.e. during the tachyonic growth. The small deviations at $k\leq
1311: m$ are simulation artifacts due to the initial UV cut-off
1312: implementation and soon disappear.}
1313: \label{PijSpec}
1314: \vspace*{-3mm}
1315: \end{figure}
1316: %%%%%%%%%%%%%%%%%%%%%%%%%
1317: 
1318: The comparison between the tensor modes $h_{ij}(k,t)$ and the
1319: numerical results is somewhat more complicated. We should first
1320: compute the effective anisotropic tensor T$_{ij}({\bf
1321: k},t)$~(\ref{GWfakeSource}) from the gradients of the Higgs field (those
1322: of the inflaton are not relevant during the tachyonic growth), as
1323: follows,
1324: \begin{equation}
1325: \indent 
1326: \tilde\Pi_{ij}({\bf k},t) = \int {d^3{\bf x}\,e^{-i{\bf k}{\bf x}}
1327: \over(2\pi)^{3/2}}
1328: \left[\nabla_i\phi^a\,\nabla_j\phi^a({\bf x},t) 
1329: %-\frac{1}{3}(\nabla\phi)^2\delta_{ij}
1330: \right]\,,
1331: \end{equation}
1332: where
1333: \begin{equation}
1334: \indent 
1335: \nabla_i\phi^a({\bf x},t) = \int {d^3{\bf q}\over(2\pi)^{3/2}}
1336: \,iq_i\,\tilde{\phi^a}({\bf q},t)\,e^{-i{\bf q}{\bf x}}\,.
1337: \end{equation}
1338: After performing the integral in ${\bf x}$ and using the delta
1339: function to eliminate ${\bf q}'$, we make a change of variables
1340: ${\bf q}\to {\bf q}+{\bf k}/2$, and integrate over ${\bf q}$,
1341: with which the Fourier transform
1342: of the anisotropic stress tensor becomes 
1343: \begin{eqnarray}\label{Pijk}
1344: \tilde\Pi_{ij}({\bf k},t) = 
1345: k_i\,k_j\,{A(\tau)\over B(\tau)\sqrt2}\,
1346: \Psi\left[{1\over2},0;{B(\tau)k^2\over4}\right]
1347: \,e^{-{1\over4}B(\tau)k^2}, \nonumber \\
1348: \end{eqnarray}
1349: which gives a very good approximation to the numerical results, see
1350: Fig.~\ref{PijSpec}, with $\Psi(1/2,0;z)\simeq(\pi^{-1}+z)^{-1/2}$ 
1351: being the Kummer function.
1352: 
1353: Finally, with the use of $\tilde\Pi_{ij}({\bf k},t)$, we can compute 
1354: the tensor fields,
1355: \begin{eqnarray}
1356: \hspace{3mm}
1357: h_{ij}({\bf k},t)&=&(16\pi G)\,\int_0^t\,dt'\,{\sin k(t-t')\over k}\,
1358: \tilde\Pi_{ij}\\
1359: \hspace{3mm}
1360: \partial_0 h_{ij}({\bf k},t)&=&(16\pi G)\,\int_0^t\,dt'\,\cos k(t-t')\,
1361: \tilde\Pi_{ij}.
1362: \end{eqnarray}
1363: Using the analytic expression in Eq.~(\ref{Pijk}) one can perform the
1364: integrals and obtain expressions that agree surprisingly well with
1365: the numerical estimates. This allows one to compute the density in
1366: gravitational waves, $\rho_{_{\rm GW}}$, at least during the initial 
1367: tachyonic stage in terms of analytical functions, and we reproduce
1368: the numerical results, see Fig.~\ref{fig4}.
1369: 
1370: We will now compare our numerical results with the analytical estimates.
1371: The tachyonic growth is dominated by the faster than exponential
1372: growth of the Higgs modes towards the true vacuum.
1373: The (traceless) anisotropic strees tensor $\Pi_{ij}$ grows rapidly to
1374: a value of order $k^2|\phi|^2 \sim 10^{-3}\,m^2v^2$, which gives a
1375: tensor perturbation 
1376: \begin{equation}
1377: \indent
1378: \left|h_{ij} h^{ij}\right|^{1/2} \sim 16\pi
1379: G v^2 (m\Delta t)^2 10^{-3}\,,
1380: \end{equation}
1381: and an energy density in GW, 
1382: \begin{equation}
1383: \indent
1384: \rho_{_{\rm GW}}/\rho_0 \sim 64\pi G v^2\,(m\Delta t)^2 10^{-6} \sim Gv^2\,, 
1385: \end{equation}
1386: for $m\Delta t \sim 16$.  In the case at hand, with $v=10^{-3}\,M_P$, we
1387: find $\rho_{_{\rm GW}}/\rho_0 \sim 10^{-6}$ at symmetry breaking, which
1388: coincides with the numerical simulations at that time, see
1389: Fig.~\ref{fig3}. 
1390: 
1391: As shown in Ref.~\cite{symmbreak}, the spinodal instabilities grow
1392: following the statistics of a gaussian random field, and therefore one
1393: can use the formalism of~\cite{BBKS} to estimate the number of peaks
1394: or lumps in the Higgs spatial distribution just before symmetry
1395: breaking.  As we will discuss in the next section, these lumps will
1396: give rise via non-linear growth to lump invagination and the formation
1397: of bubble-like structures with large density gradients, expanding at
1398: the speed of light and colliding among themselves giving rise to a
1399: large GWB. The size of the bubbles upon collision is essentially
1400: determined by the distance between peaks at the time of symmetry
1401: breaking, but this can be computed directly from the analysis of
1402: gaussian random fields, as performed in Ref.~\cite{symmbreak}.
1403: 
1404: This analysis works only for the initial (linear) stage before symmetry
1405: breaking.  Nevertheless, we expect the results to extrapolate to later
1406: times since once a bubble is formed around a peak, it remains there at
1407: a fixed distance from other bubbles. This will give us an idea of the
1408: size of the bubbles at the time of collision.
1409: 
1410: We estimate the number density of peaks as~\cite{BBKS}
1411: \begin{equation}
1412: \indent n_{\rm peak}(\tau) = {2\over3\sqrt3\,\pi^2}\xi_0(\tau)^{-3}(\nu^2-1)
1413: \exp[-\nu^2/2]\,, 
1414: \end{equation}
1415: where $\nu=\phi_c/\sigma(\tau)$ corresponds to the ratio of the field
1416: threshold $\phi_c$ over the dispersion
1417: \begin{equation}
1418: \indent \sigma(\tau) = {\sqrt\lambda\over\pi}(2V)^{1/3}\Big({A(\tau)\over 
1419: B(\tau)}\Big)^{1/2}\,,
1420: \end{equation}
1421: with $A(\tau)$ and $B(\tau)$ given in Eq.~(\ref{ABT}). The average
1422: size of the gaussian lumps is $\xi_0(\tau)=2B^{1/2}(\tau)\,m^{-1}$,
1423: where time is given in units $\tau = (2V)^{1/3}\,mt$, see
1424: Ref.~\cite{symmbreak}
1425: 
1426: The distance between peaks can be estimated as twice the radius of the
1427: average bubble, with volume $V_{\rm peak} = 4\pi/3\,R_{\rm peak}^3$.
1428: Since the total volume $L^3$ is divided into $N_{\rm peak}$ bubbles,
1429: we find 
1430: \begin{equation}
1431: \indent d_{\rm peak} = 2R_{\rm peak} = {1\over ma}\,\Big({6\over\pi
1432: n_{\rm peak}}\Big)^{1/3}\,a,
1433: \end{equation}
1434: which is typically of order 30 to 40 lattice units, for $\phi_c \simeq
1435: 0.5 - 0.8$, $V=0.024$ and $\lambda=0.125$, with lattices sizes given
1436: by $p_{\rm min}=0.15\,m$ and $N=128$.
1437: 
1438: What is interesting is that decreasing either $\lambda$ or $V$, the
1439: distance between initial lumps increases and thus also the size of the
1440: final bubbles upon collision. As we will show in the next section, the
1441: amplitude of GW depends on the bubble size squared, and therefore it
1442: is expected that for lower lambda we should have larger GW amplitude.
1443: We have not seen, however, such an increase in amplitude, but a 
1444: detailed analysis is underway and will be presented elsewhere.
1445: 
1446: \subsection{Bubble collisions}
1447: 
1448: The production of gravitational waves in the next stage proceeds
1449: through `bubble' collisions. In Ref.~\cite{tachyonic} we showed
1450: explicitly that symmetry breaking is not at all a homogeneous
1451: process. During the breaking of the symmetry, the Higgs field develops
1452: lumps whose peaks grow up to a maximum value $|\phi|_{\rm max}/v =
1453: 4/3$ (see paper I), and then decrease creating approximately
1454: spherically symmetric bubbles, with ridges that remain above $|\phi| =
1455: v$. Finally, neighboring bubbles collide and the symmetry gets further
1456: broken through the generation of higher momentum modes.  Since
1457: initially only the Higgs field sources the anisotropic stress-tensor
1458: $\Pi_{ij}$, then we expect the formation of structures (see section
1459: IV.A) in the tensor metric perturbation, correlated with the Higgs
1460: lumps. The dependence of the $h_{ij}$ tensor on the gradient of the
1461: Higgs field, see~Eq.(\ref{GWeq}), is responsible of the formation of
1462: those structures in the energy density spatial distribution of the
1463: GWB.
1464: 
1465: In section V of this paper we will give account of the explicit form
1466: of the structures developed in the spatial distribution of
1467: $\rho_{_{\rm GW}}$ related with the first collisions among the
1468: bubble-like structures of the Higgs field. We will present
1469: simultaneously the evolution of both the Higgs' spatial distribution
1470: when the first bubbles start colliding, and of the corresponding
1471: structures in the GW energy density $\rho_{_{\rm GW}}$. We leave for a
1472: forthcoming publication the details of an analytical formalism
1473: describing the formation and subsequent evolution of such GW
1474: structures. In this sub-section we will just give an estimate of the
1475: burst in GW produced by the first collisions of the Higgs bubble-like
1476: structures.
1477: 
1478: Thus, as described in Ref.~\cite{KosowskyTurner} for the case of first
1479: order phase transitions, which involves the collision of vacuum
1480: bubbles, we can give a simple estimate of the order of magnitude of
1481: the energy fraction radiated in the form of gravitational waves when
1482: two Higgs bubble-like structures collide. In general, the problem of
1483: two colliding bubbles has several time and length scales: the duration
1484: of the collision, $\Delta t$; the bubbles' radius $R$ at the moment of
1485: the collision; and the relative speed of the bubble walls. In section
1486: IV.B we found that the typical size of bubbles upon collisions, is of
1487: the order of $R\approx 10m^{-1}$, while the growth of the bubble's
1488: wall is relativistic. Then we can assume than the time scale
1489: associated with bubble collisions is also $\Delta t\,\sim\,R$.
1490: Assuming the bubble walls contain most of the energy density, and
1491: since they travel close to the speed of light, see paper I, it is 
1492: expected that the asymmetric collisions will copiously produce GW.
1493: 
1494: Far from a source that produces gravitational radiation, the dominat
1495: contribution to the amplitude of GW is given by the acceleration of
1496: the quadrupole moment of the Higgs field distribution. Given
1497: the energy density of the Higgs field, $\rho_{\rm H}$, we can compute
1498: the (reduced) quadrupole moment of the Higgs field spatial
1499: distribution, $Q_{ij} = \int d^3x\,(x_i x_j -
1500: x^2\delta_{ij}/3)\,\rho_{\rm H}(x)$, such that the amplitude of the
1501: gravitational radiation, in the TT gauge, is given by 
1502: $h_{ij}\,\sim\,(2G/r)\ddot Q_{ij}$. A significant amount of energy can be
1503: emitted in the form of gravitational radiation whenever the
1504: quadrupole moment changes significantly fast: through the bubble
1505: collisions in this case. The power carried by these waves can be
1506: obtained via~(\ref{rhoTotal}) as
1507: \begin{eqnarray}
1508: \indent P_{_{\rm GW}} = \frac{G}{8\pi}\int d\Omega\,\left\langle  
1509: \dddot{Q}_{ij}\dddot{Q}^{ij}\right\rangle\,. 
1510: \end{eqnarray} 
1511: Omitting indices for simplicity, as the power emitted in gravitational
1512: waves in the quadrupole approximation is of order $P_{_{\rm
1513: GW}}\,\sim\,G(\dddot{Q})^2$, while the quadrupole moment is of order
1514: $Q \sim R^5\rho_{\rm H}$, we can estimate the power emitted in GW
1515: upon the collision of two Higgs bubbles as
1516: \begin{equation}
1517: \indent P_{_{\rm GW}}\,\sim\,G\left(\frac{R^5\rho}{R^3}\right)^2\,
1518: \sim\,G\rho_{\rm H}^2\,R^4
1519: \end{equation} 
1520: The fraction of energy density carried by these waves, $\rho_{_{\rm
1521: GW}}\,\sim\,P_{_{\rm GW}}\Delta t/R^3\,\sim\,P_{_{\rm
1522: GW}}/R^2\,\sim\,G\rho_{\rm H}^2\,R^2$, compared to that of the initial
1523: energy stored in the two bubble-like structures of the Higgs field,
1524: will be $\rho_{_{\rm GW}}/{\rho_{\rm H}} = G\rho_{\rm H}R^2$.  Since
1525: the expansion of the universe is negligible during the bubble
1526: collision stage, the energy that drives inflaton, $\rho_0\,\sim\,
1527: m^2v^2$, is transferred essentially to the Higgs modes during
1528: preheating, within an order of magnitude, see Fig.~\ref{fig3}. Thus,
1529: recalling that $R\,\sim 10m^{-1}$, the total fraction of energy in GW
1530: produced during the bubble collisions to that stored in the Higgs
1531: lumps formed at symmetry breaking, is given by
1532: \begin{eqnarray}\label{GWbubbles}
1533: \indent \frac{\rho_{_{\rm GW}}}{\rho_0}\,\sim 0.1\,G\rho_0\,R^2\,
1534: \sim\,(v/M_p)^2\,,
1535: \end{eqnarray} 
1536: giving an amplitude which is of the same order as is observed in the
1537: numerical simulations, see Fig.~\ref{fig3}.  Of course, an exhaustive
1538: analytical treatment of the production of GW during this stage of
1539: bubble collisions remains to be done, but we leave it for a future
1540: publication.
1541: 
1542: %%%%%%%%%%%%%%%%%%%%%%%%%
1543: \begin{figure}[t]%[t]
1544: %\vspace{5mm}
1545: \begin{center}
1546: %\includegraphics[width=4.cm,height=7.cm,angle=-90]{meanFields.eps}
1547: \includegraphics[width=5.5cm,height=8.5cm,angle=-90]{variance_hyb4Fld_wl_noErrors.eps}
1548: \end{center}
1549: \vspace*{-5mm}
1550: \caption{Variance of the Inflaton and the Higgs field as a function of
1551: time, the former normalized to its critical value, the latter
1552: normalized to its \textit{v.e.v.}. As expected in a turbulent regime,
1553: these variances follow a power law $\sim t^{-2p}$ with $p$ a certain
1554: critical exponent, although the slope of the Inflaton's variances
1555: evolves in time. The curves are produced from an average over 10 
1556: different statistical realizations.}
1557: \label{variances_fig}
1558: \vspace*{-3mm}
1559: \end{figure}
1560: %%%%%%%%%%%%%%%%%%%%%%%%%
1561: 
1562: \subsection{Turbulence}
1563: 
1564: The development of a turbulent stage is expected from the point of
1565: view of classical fields, as turbulence usually appears whenever there
1566: exists an active (stationary) source of energy localized at some scale
1567: $k_{\rm in}$ in Fourier space. As first pointed out by
1568: Ref.~\cite{MichaTkachev}, in reheating scenarios the coherently
1569: oscillating inflaton zero-mode plays the role of the pumping-energy
1570: source, acting at a well defined scale $k_{\rm in}$ in Fourier space,
1571: given by the frequency of the inflaton oscillations. Thus, the
1572: inflaton zero-mode pumps energy into the rest of the fields that
1573: couple to it as well as into the non-zero modes of the inflaton field
1574: itself. Apart from $k_{\rm in}$, there is no other scale in Fourier
1575: space where energy is accummulated, dissipated and/or infused. So, as
1576: turbulence is characterized by the transport of some conserved
1577: quantity, energy in our case, we should expect a flow of energy from
1578: $k_{\rm in}$ towards higher (direct cascade) or smaller (inverse
1579: cascade) momentum modes. In typical turbulent regimes of classical
1580: fluids, there exits a sink in Fourier space, corresponding to that
1581: scale at which the (direct) cascade stops and energy gets
1582: dissipated. However, in our problem there is no such sink so that the
1583: transported energy cannot be dissipated, but instead it is used to
1584: populate high-momentum modes. For the problem at hand, there exists a
1585: natural initial cut-off $k_{\rm out} \sim \lambda^{1/2}v,$ such that
1586: only long wave modes within $k<k_{\rm out}$, develop the spinodal
1587: instability. Eventually, after the tachyonic growth has ended and the
1588: first Higgs' bubble-like structures have collided, the turbulent
1589: regime is established. Then the energy flows from small to greater
1590: scales in Fourier space, which translates into the increase of $k_{\rm
1591: out}$ in time.
1592: 
1593: %%%%%%%%%%%%%%%%%%%%%%%%%
1594: \begin{figure}[t]%[t]
1595: %\vspace{5mm}
1596: \begin{center}
1597: %\includegraphics[width=4.cm,height=7.cm,angle=-90]{meanFields.eps}
1598: \includegraphics[width=5.5cm,height=8.5cm,angle=-90]{n_Higgs_p01_Evol+Scal.eps}
1599: \end{center}
1600: \vspace*{-5mm}
1601: \caption{Some snapshots of the evolution of the spectral particle
1602: occupation numbers of the Higgs field at different times, each
1603: averaged over 10 statistical realizations. We multiply them by $k^4$
1604: so we can see better the scaling behaviour. In the upper right corner,
1605: we plot the inverse relation of~(\ref{selfSimilar}), $n_0(k t^{-p})
1606: = t^{\gamma p}n(k,t)$, also averaged over 10 realizations for each
1607: time. The scaling behaviour predicted by wave kinetic turbulent
1608: theory~\cite{MichaTkachev}, is clearly verified.}
1609: \label{higgsScaling_fig}
1610: \vspace*{-3mm}
1611: \end{figure}
1612: %%%%%%%%%%%%%%%%%%%%%%%%%
1613: 
1614: %%%%%%%%%%%%%%%%%%%%%%%%%
1615: \begin{figure}[t]%[t]
1616: %\vspace{5mm}
1617: \begin{center}
1618: \includegraphics[width=5.5cm,height=8.5cm,angle=-90]{n_Inflaton_p0175_Evol+Scal.eps}
1619: \end{center}
1620: \vspace*{-5mm}
1621: \caption{Different times of the evolution of the particle occupation
1622: numbers spectra of the Inflaton, multiplied by $k^4$, and averaged
1623: over 10 statistical realizations or each time. Again, in the upper
1624: right corner, we plot the inverse relation of~(\ref{selfSimilar}),
1625: $n_0(k) = t^{\gamma p}n(kt^{p},t)$, also averaged over 10
1626: realizations for each time.}
1627: \label{inflatonScaling_fig}
1628: \vspace*{-3mm}
1629: \end{figure}
1630: %%%%%%%%%%%%%%%%%%%%%%%%%
1631: 
1632: In Ref.~\cite{magnetic} we already accounted for the turbulent stage
1633: reached in a hybrid model with gauge fields. However, we don't
1634: consider gauge fields here, so the number of degrees of freedom is
1635: different from that of Ref.~\cite{magnetic} and, therefore, the turbulent
1636: dynamics of the Inflaton and the Higgs fields should be different. In
1637: particular, when the turbulence has been fully established, if the
1638: wave (kinetic) turbulence regime of the fields' dynamics is valid, the
1639: time evolution of the variance of a turbulent field $f({\bf x},t)$,
1640: should follow a power-law-like scaling~\cite{MichaTkachev}
1641: \begin{equation}
1642: \label{variances}
1643: \indent \mathrm{Var}(f(t)) = \left\langle f(t)^2\right\rangle -
1644: \left\langle f(t)\right\rangle^2 \propto \ t^{-2p} \,,
1645: \end{equation}
1646: with $p = 1/(2N-1)$ and $N$ the number of scattering fields in a
1647: `point-like collision'. In fact, such time behaviour corresponds only
1648: to the case of the so called \textit{free turbulence}, when the energy
1649: stored in the pumping source is subdominant to the energy in the
1650: turbulent fields. In our case, this condition is reached very soon
1651: after the symmetry breaking, so we don't expect a significant stage of
1652: \textit{driven turbulence}, which would make the variance to increase
1653: (Only the inflaton seems to increase its variance between $mt=10$ and
1654: $mt=30$, but it is not very pronounced). In Fig.~\ref{variances_fig}
1655: we have plotted the time evolution of the variances of the Inflaton
1656: $\chi$ and of the Higgs modulus $\phi = \sqrt{\sum_a\phi_a^2}$, and
1657: fitted the data with a power-law like~(\ref{variances}), obtaining
1658: \begin{flushleft}
1659: \begin{tabular}{rrll}
1660: \vspace{.2cm} \,\,\,\, & Inflaton: & $p^{-1}_{_I} = 5.1 \pm 0.2$, &
1661: \,\,\,[35:85] \\ \vspace{.2cm} \,\,\,\,& Inflaton: & $p^{-1}_{_I} =
1662: 9.03 \pm 0.03$, & \,\,\,[350:2000] \\ \vspace{.1cm} \,\,\,\, & Higgs:
1663: & $p^{-1}_{_H} = 7.02 \pm 0.01$, & \,\,\,[50:2000]
1664: \end{tabular}
1665: \end{flushleft}
1666: where the last brackets on the right correspond to the range in time 
1667: (in units of $m^{-1}$) for which we fitted the data. As can be seen in
1668: Fig.~\ref{variances_fig}, the slope of the Higgs field (in
1669: logarithmic scale), $2p_{_H}\,\sim\,2/7$, remains approximately
1670: constant in time, corresponding to a 4-field dominant interaction.
1671: However, the slope of the Inflaton's variance increases in time, 
1672: i.e. the critical exponent $p_{_I}$ of the Inflaton decreases, 
1673: until it reaches a stationary stage at $mt\sim100$. 
1674: Since $p_{_I}$ is related to the number $N$ of fields interacting in a
1675: collision, if there was a change from one dominant multi-field
1676: interaction to another, this should produce a time-dependent effective
1677: $p_{_I}$, as seen in Fig.~\ref{variances_fig}. However, we will not
1678: try to explain here the origin of such an effective critical exponents
1679: as extracted from the simulations. We will just stress that we 
1680: have checked the robustness of those values under different lattice 
1681: configurations ($N,p_{\rm min}$) and different statistical realizations, 
1682: discarding this way a possible lattice artefact effect. 
1683: As we will see, the critical exponents $p$ determines the speed with 
1684: which the turbulent particle distribution moves over momentum space, 
1685: so this is a crucial parameter. Moreover, although both the classical 
1686: modes of the Inflaton
1687: and the Higgs contribute to the production of GW, the Inflaton's
1688: occupation numbers decrease faster than those of the Higgs so, after a
1689: given time, only the Higgs' modes remain as the main source of GW.
1690: 
1691: Actually, when turbulence is developed, it is expected that the
1692: distribution function of the classical turbulent fields, the inflaton
1693: and the Higgs here, follow a self-similar
1694: evolution~\cite{MichaTkachev}
1695: \begin{eqnarray}
1696: \label{selfSimilar}
1697: \indent n(k,t) = t^{-\gamma\,p}n_0(k\,t^{-p})
1698: \end{eqnarray}
1699: with $p$ the critical exponent of the fields' variances and $\gamma$ a
1700: certain factor $\sim O(1)$, which depends on the type of turbulence
1701: developed. It is precisely this way that the exponent $p$ determines
1702: the speed of the particles' distribution in momentum space: given a
1703: specific scale $k_c$ such that, for example, the occupation number has
1704: a maximum, that scale evolves in time as $k_c(t) = k_c(t_0)(t/t_0)^{p}$. 
1705: We have seen that the evolution of the Higgs occupation number follows
1706: Eq.~(\ref{selfSimilar}) with $p \approx 1/7$, as expected from the
1707: Higgs variance, and $\gamma \approx 2.7$. Whereas the evolution of the
1708: Inflaton occupation number follows~(\ref{selfSimilar}) even more
1709: accurately than the Higgs, with an ``effective" exponent $p \approx
1710: 1/5$, and $\gamma \approx 3.9$. Since the slope of the inflaton's
1711: variance changes in time, the value of the exponents of the inflaton's 
1712: scaling relation will require further investigation. However, despite 
1713: this time evolution of the Inflaton variance, Eq.~(\ref{selfSimilar})
1714: is very well fulfilled by the Inflaton with the given effective 
1715: exponents. So we can perfectly obtain the universal $n_0(k)$ function 
1716: for the Inflaton as well as for the Higgs.
1717: 
1718: In Figs.~\ref{higgsScaling_fig} and~\ref{inflatonScaling_fig} we have
1719: plotted the occupation numbers of the Higgs and the Inflaton, also
1720: inverting the relation of Eq.~(\ref{selfSimilar}) in order to extract
1721: the \textit{universal} time-independent $n_0(k)$ functions of each
1722: field. As shown in those figures, the distributions follow the
1723: expected scaling behaviour. However, for the range of interest of $k$,
1724: there are small discrepancies of order 0.1-4\% (depending on $k$)
1725: among the different universal functions $n_0^{(i)}(k)$, as obtained 
1726: inverting Eq.~(\ref{selfSimilar}) at different times $mt_i$. The universal
1727: functions $n_0(k)$ plotted in Figs.~\ref{higgsScaling_fig} 
1728: and~\ref{inflatonScaling_fig} have been obtained from averaging over
1729: ten statistical realizations for each time.
1730: 
1731: %%%%%%%%%%%%%%%%%%%%%%%%%
1732: \begin{figure}[t]%[t]
1733: %\vspace{5mm}
1734: \begin{center}
1735: \includegraphics[width=5.5cm,height=8.5cm,angle=-90]{GW_t1000.eps}
1736: \end{center}
1737: \vspace*{-5mm}
1738: \caption{Time evolution of the GW spectra from mt = 6 to mt = 2000. The amplitude of the spectra seems to saturate after mt $\sim 100$, although the high momentum tail still moves slowly to higher values of $k$ during the turbulent stage.}
1739: \label{fig10}
1740: \vspace*{-3mm}
1741: \end{figure}
1742: %%%%%%%%%%%%%%%%%%%%%%%%%
1743: 
1744: The advantage of the development of a turbulence behaviour is obvious:
1745: it allows us to extrapolate the time evolution of the fields'
1746: distributions till later times beyond the one we can reach with the
1747: simulations. Moreover, the fact that the turbulence develops so early
1748: after the tachyonic instability, also allow us to check for a long
1749: time of the simulation, the goodness of the description of the
1750: dynamics of the fields, given by the turbulent kinetic theory
1751: developed in Ref.~\cite{MichaTkachev}. We have fitted the averaged
1752: universal functions $n_0(k)$ with expressions of the form $k^4\,n_0(k)
1753: = P(k)e^{-Q(k)}$, with $P(k)$ and $Q(k)$ polynomials in $k$, giving: 
1754: \begin{eqnarray}
1755: \begin{array}{lll}
1756: \label{Inflaton_universal}
1757: {\rm Inflaton:} & P(k) = 486.2k^3 & Q(k) = 6.39k \\
1758: \,&\,& \\
1759: \label{Higgs_universal}
1760: {\rm Higgs:} & P(k) = 2.96k^3 & Q(k) = 2.26k^2 - 3.18k
1761: \end{array}
1762: \end{eqnarray} 
1763: There is no fundamental meaning for these expressions, but it is very
1764: useful to have such an analytical control over $n_0(k)$, since this
1765: allows us to track the time-evolution of $n(k,t)$ through
1766: Eq.~(\ref{selfSimilar}). Actually, the classical regime of the
1767: evolution of some bosonic fields ends when the system can be relaxed
1768: to the Bose-Einstein distribution. We are now going to estimate the
1769: moment in which the initial energy density gets fully transferred to
1770: the Higgs classical modes.  Using Eq.(\ref{selfSimilar}) and the
1771: fit~(\ref{Higgs_universal}), we find that the initial energy density 
1772: % for $n(k)\,\sim \,1$ 
1773: is totally transfered to the Higgs when (in units $m=1$)
1774: \begin{equation}
1775: \indent
1776: \rho_0 = {1\over4\lambda} =
1777: \int {dk\over k}\,{k^3\over2\pi^2}\,k\,n(k,t) =
1778: {7.565\over2\pi^2}\, t^{(4-\gamma)p}\,,
1779: \end{equation}
1780: where we have assumed that the Higgs' modes have energy $E_k(k,t) =
1781: k\,n(k,t)$. In our case, with $\lambda=1/8$, the conversion of the
1782: initial energy density into Higgs particles and therefore into
1783: radiation is complete by $mt \sim 6\times10^4$. Therefore, if we 
1784: consider this value as a lower bound for the time that classical 
1785: turbulence requires to end, we see that turbulence last for a very 
1786: long time compared to the time-scale of the initial tachyonic and bubbly 
1787: stages. Thus, if GW were significatively sourced during turbulence, 
1788: one should take into account corrections from the expansion of the universe.
1789: 
1790: In Fig.~\ref{fig10}, we show the evolution of the GW spectra up to
1791: times mt $= 2000$, for a lattice of (N,$p_{\rm min}$) = (128,.15). It
1792: is clear from that figure that the amplitude of the GW saturates to a
1793: value of order $\rho_{gw}/\rho_0 \approx 2\cdot10^{-6}$. At mt
1794: $\approx 50$, the maximum amplitude of the spectra has already reached
1795: $\rho_{gw}/\rho_0 \approx 10^{-6}$, while at time mt $\approx 100$,
1796: the maximum has only grown a factor of 2 with respect to mt $\approx
1797: 50$. From times mt $\approx 150$ till the maximum time we reached in
1798: the simulations, mt = 2000, the maximum of the amplitude of the
1799: spectrum does not seem to change significantly, slowly increasing from
1800: $\approx 2\cdot10^{-6}$ to $\approx 2.5\cdot10^{-6}$. Despite this
1801: saturation, we see in the simulations that the the long momentum tail
1802: of the spectrum keeps moving towards greater values. This displacement
1803: is precisely what one would expect from turbulence, although it is
1804: clear that the amplitude of the new high momentum modes never exceed
1805: that of lower momentum. In order to disscard that this displacement
1806: towards the UV is not a numerical artefact, one should further
1807: investigate the role played by the turbulent scalar fields as a source
1808: of GW. Here, we just want to remark that the turbulent motions of the
1809: scalar fields, seem not to increase significatively anymore the total
1810: amplitude of the GW spectrum. Indeed, in a recent paper
1811: \cite{DufauxGW} where GW production at reheating is also considered,
1812: it is stated that GW production from turbulent motion of classical
1813: scalar fields, should be very supressed. That is apparently what we
1814: observe in our simulations although, as pointed above, this issue
1815: should be investigated in a more detailed way. Anyway, here we can
1816: conclude that the expansion of the Universe during reheating in these
1817: hybrid models, does not play an important role during the time of GW
1818: production, and therefore we can be safely ignore it.
1819: 
1820: %\vspace{1cm}
1821: 
1822: \begin{widetext}
1823: 
1824: \begin{figure}[t]
1825: \centering
1826: \includegraphics[width=7.5cm,height=14.5cm,angle=270]{ring.eps}
1827: \caption{Model: $\lambda = 10^{-3}$, $g^2=1$. In the left picture we
1828: show a spatial section of $\left<\phi\right>$. We can see how a
1829: spherical lump is growhing. In the right picture we can see the
1830: structure of the $\rho_{GW}$ in the same place.  A ring is forming
1831: around the Higgs lump. More complex structures are formed in the
1832: regions in which the Higgs bubbles are next, and the GW grow in the
1833: boundary of this lumps, where the gradient of the Higgs and therefore
1834: the $\Pi_{ij}$ tensor grow in this region.}
1835: \label{fig:ring1}
1836: \end{figure}
1837: 
1838: \end{widetext}
1839: 
1840: \section{Spatial sections and local GW production}
1841: 
1842: \begin{figure}[htb]
1843: \vspace{5mm}
1844: \centerline{
1845: \subfigure{
1846: \includegraphics[width=4.5cm,height=9.0cm,angle=-90]{expansion_t10.eps}}}
1847: \centerline{
1848: \subfigure{
1849: \includegraphics[width=4.5cm,height=9.0cm,angle=-90]{expansion_t11.eps}}}
1850: \caption{Here we have got the time evolution of the previous ring
1851: (picture \ref{fig:ring1}) near the symmetry breaking. The bubble is
1852: growing (mt = 10-11), until the symmetry breaking time (mt = 12).}
1853: \label{fig:bubble_expansion1}
1854: \end{figure}
1855: 
1856: 
1857: \begin{figure}[htb]
1858: \vspace{5mm}
1859: \centerline{
1860: \subfigure{
1861: \includegraphics[width=4.5cm,height=9.0cm,angle=-90]{expansion_t12.eps}}}
1862: \centerline{
1863: \subfigure{
1864: \includegraphics[width=4.5cm,height=9.0cm,angle=-90]{expansion_t13.eps}}}
1865: \caption{The Higgs lump begins to invaginate, and the GW ring
1866: expands (mt = 12-13). A similar behavior is observed in a smaller
1867: lump below the biggest Higgs lump, in the same pictures.}
1868: \label{fig:bubble_expansion2}
1869: \end{figure}
1870: 
1871: 
1872: In this section, we show a sequence of snapshots ($mt = 5-20$) of the
1873: evolution of the spatial distribution, before the fields are driven to
1874: the turbulent stage. We find that the first stages of the GW dynamics
1875: is strongly correlated with the dynamics of the Higgs oscillations
1876: that give rise to symmetry breaking. A qualitative way of
1877: understanding this question is to analyse the spatial structure of the
1878: $\Pi_{ij}$ tensor, built from spatial gradients of the Higgs and
1879: inflaton fields.  Since the oscillations of $\langle\phi\rangle$ are
1880: due to rapid changes of the Higgs' values in its way of symmetry
1881: breaking, this induces great variations in the behaviour of the
1882: spatial gradients. We are now going to analyse briefly the different
1883: stages showing the most representatives images. Shortly, it will be
1884: available in our web page, a bigger selection of pictures and movies
1885: \cite{LATTICEWEB}.
1886: 
1887: An interesting conclusion from the set of Figs.~\ref{fig:ring1} $-$
1888: \ref{fig:compression2} is that the Higgs evolution from lump growth,
1889: through invagination to bubble collisions, has a direct translation
1890: into the corresponding growth of gravitational wave energy density.
1891: Not only does the volume-averaged amplitude $\rho_{\rm GW}$ follow
1892: the Higgs time evolution, but the individual local features in the
1893: GWB seem to correspond very closely with the Higgs features.
1894: 
1895: In the first stage both Higgs and GW backgrounds grow very fast. The
1896: lumps which grow in the Higgs background induce structures around
1897: these, through the gradients appearing in the $\Pi_{ij}$ tensor.  The
1898: geometry of the gravitational structures comes from the position of
1899: the Higgs lump. A typical structure for an isolated lump is a ring of
1900: gravitational waves, see Fig.~\ref{fig:ring1}. More complex structures
1901: can be formed around the position of the Higgs lumps. Before symmetry
1902: breaking these lumps grow according to the previous analysis,
1903: generating domains with a great density of gravitational energy.
1904: 
1905: 
1906: \begin{figure}[t]
1907: \vspace{5mm}
1908: \centerline{
1909: \subfigure{\includegraphics[width=4.5cm,height=9cm,angle=-90]{rupt_l_0.125_g2_0.5_t13.0.eps}}}
1910: \centerline{
1911: \subfigure{\includegraphics[width=4.5cm,height=9cm,angle=-90]{rupt_l_0.125_g2_0.5_t13.5.eps}}}
1912: \caption{After symmetry breaking the expansion of Higgs lump
1913: compresses the GW, until the Higgs gradient changes in the first
1914: oscillation (mt =13-13.5).}
1915: \label{fig:compression1}
1916: \end{figure}
1917: 
1918: % \end{widetext}
1919: 
1920: %\begin{widetext}
1921: 
1922: \begin{figure}
1923: \vspace{5mm}
1924: \centerline{
1925: \subfigure{\includegraphics[width=4.5cm,height=9cm,angle=-90]{rupt_l_0.125_g2_0.5_t14.0.eps}}}
1926: \centerline{
1927: \subfigure{\includegraphics[width=4.5cm,height=9cm,angle=-90]{rupt_l_0.125_g2_0.5_t14.5.eps}}}
1928: \caption{At this moment, when the Higgs falls, the
1929: GW strutrure is divided in two waves (mt = 14). These wave fronts
1930: propagates in opposite directions (mt = 14.5).}
1931: \label{fig:compression2}
1932: \end{figure}
1933: 
1934: %\end{widetext}
1935: 
1936: 
1937: The second stage begins when $\langle\phi\rangle = v$ and the symmetry
1938: breaking starts, then the Higgs lumps invaginate and expand, producing
1939: the growth of gravitational waves around of these structures, see
1940: Figs. \ref{fig:bubble_expansion1} and \ref{fig:bubble_expansion2}, one
1941: can see that whenever the bubble walls expand, the variation in the
1942: gradient of the Higgs' field induces the expansion of the GW ring. In
1943: the case of the rings, if the lump is very isolated, the expansion
1944: induces the ring to dilute and disappear, by Gauss law.  In practice,
1945: however, the lumps are never isolated and bubbles collide before the
1946: gradients (and thus the GW) die away.
1947: 
1948: In the case when two Higgs' bubble-like structures are close by, the
1949: expansion of their walls compresses the GW structures.  This expansion
1950: continues until the first Higgs oscillation, see Fig.~\ref{fig4}. If
1951: the distance between Higgs' structures is small, then the GW can be
1952: diluted, whereas in the other case, a remnant string-like GW structure
1953: survives, and when the Higgs background goes to zero this GW structure
1954: becomes divided into two waves that propagate in opposite directions,
1955: as one can see in Figs.~\ref{fig:compression1} and
1956: \ref{fig:compression2}, which show four snapshots of this process.  A
1957: similar behaviour is observed in the second oscillations.
1958: 
1959: Finally, the wave fronts propagate, colliding among themselves,
1960: driving the system to the stage of turbulence. We will leave for a
1961: future publication the detailed analysis of the GW production at the
1962: bubble collisions and the subsequent turbulent period.
1963: %after preheating in hybrid models.
1964: 
1965: %\vspace{1cm}
1966: 
1967: \section{Gravitational Waves from Chaotic Inflation}
1968: 
1969: The production of a relic GWB at reheating was first 
1970: addressed by Khlebnikov and Tkachev (KT) in 
1971: Ref.~\cite{TkachevGW}, both for the quadratic and quartic chaotic
1972: inflation scenarios. In these models, the long-wavelength part of the
1973: spectrum is dominated by the gravitational bremsstrahlung associated
1974: with the scattering of the extra scalar particles off the inflaton
1975: condensate, `evaporating' this way the inflaton particles. Using
1976: this fact, KT estimated analytically the amplitude of the power
1977: spectra of GW for the low frequency end of the spectrum,
1978: corresponding to wavelengths of order the size of the horizon at
1979: rescattering. Moreover, KT also studied the GW power spectra
1980: numerically, although just for the massless inflaton case. Recently,
1981: chaotic scenarios were revisited in Ref.~\cite{EastherLim,EastherLim2}, 
1982: accompanied by more precise numerical simulations at different energy 
1983: scales, including the case of a massive inflaton. Finally, also very 
1984: recently, Ref.~\cite{DufauxGW} studied in a very detailed way, 
1985: both analitical and numerically, the evolution of GW produced 
1986: at preheating in the case of a massless inflaton with an extra scalar field.
1987: 
1988: In Refs.~\cite{TkachevGW} and~\cite{EastherLim}, the procedure
1989: to compute the GW from reheating relied on Weinberg's 
1990: formula for the energy carried by a weak gravitational radiative field
1991: in flat space-time~\cite{Weinberg}. However, in chaotic models, the
1992: expansion of the universe can not be neglected during reheating, 
1993: so Weinberg's formula can only be used in an approximated way, 
1994: if the evolution of the universe
1995: is considered as an adiabatic sequence of stationary
1996: universes. Rescaling fields by a conformal transformation, their
1997: evolution equations can be solved with a numerical integrator, while
1998: the evolution of the scale factor can be calculated
1999: analytically. Discretizing the time, the physical variables can be
2000: recovered from the conformal ones in each time step, thus allowing to
2001: compute the energy of gravitational waves in terms of the physical
2002: fields. In this paper, however, we adopt another approach\footnote{Note that 
2003: Refs.~\cite{EastherLim2} and~\cite{DufauxGW} also work 
2004: in the same theoretical framework, considering the TT tensor 
2005: perturbations on top of a flat FRW space. However, we use a 
2006: different way to extract numerically the GW power spectra, relying on the 
2007: \textit{conmutating procedure}, as detailed
2008: explained in subsection III.B} that 
2009: takes into account expansion of the universe in a
2010: self-consistent manner, and let us calculate at any time 
2011: the energy density and power spectra of the GW produced 
2012: at reheating. As explained in section III and applied to the case of 
2013: hybrid inflation in sections IV and V,
2014: we just solve numerically Eq.~(\ref{GWfakeEq}), together with those 
2015: eqs. of the other Bose
2016: fields and the scale factor, Eqs.~(\ref{inflatonEq}),(\ref{scalarEq}) and 
2017: Eqs.(\ref{hubbleDotEq}),(\ref{hubbleEq}).
2018: Then, using the projector~(\ref{projector}) into the (Fourier transformed) 
2019: solution of Eq.(\ref{GWfakeEq}), we recover the TT d.o.f corresponding 
2020: to GW. This way, we can monitor the total energy
2021: density in GW using Eq.~(\ref{rhoTotal}), or track the evolution 
2022: of the power
2023: spectrum. Using this technique, we will show in this
2024: section that we reproduce, for specific chaotic models, 
2025: similar results to those of other authors.
2026: 
2027: We adapted the publicly
2028: available LATTICEEASY code~\cite{LatticeEasy}, taking advantage of the
2029: structure of the code itself, incorparating the evolution of 
2030: Eq.~(\ref{GWeq}), together with the equations of the scalar fields, 
2031: Eqs.~(\ref{inflatonEq}) and (\ref{scalarEq}),
2032: into the staggered leapfrog integrator routine. This way, we can solve
2033: at the same time the dynamics of the scalar and tensor fields, within
2034: the framework of an expanding FRW universe~Eqs.(\ref{hubbleDotEq}) and
2035: (\ref{hubbleEq}). 
2036: 
2037: In particular, we will concentrate only in an scenario with a massless 
2038: inflaton $\chi$, either accompanied or not by an extra scalar field $\phi$.
2039: In the following, we 
2040: will describe the numerical results for GW production at reheating in such 
2041: scenarios, described by the potential
2042: \begin{equation}
2043: \indent V(\chi,\phi) = {\lambda\over 4}\chi^4 + {1\over 2}g^2\chi^2\phi^2
2044: \end{equation}
2045: Rescaling the time by
2046: \begin{eqnarray}
2047: \indent d\tau = {a(\tau)\over a(0)}\chi(0)\sqrt{\lambda}\,dt\,,
2048: \end{eqnarray}
2049: and the physical fields by a conformal transformation as
2050: \begin{eqnarray}
2051: && \chi_c(\tau) = {a(\tau)\over a(0)}{\chi(\tau)\over\chi(0)}\,, \\
2052: %\,\,\,\,\,
2053: &&\phi_c(\tau) = {a(\tau)\over a(0)}{\phi(\tau)\over\chi(0)}% \\[2mm] &&
2054: \end{eqnarray}
2055: then the equations of motion of the inflaton and of the extra scalar field,
2056: Eqs.~(\ref{inflatonEq}) and (\ref{scalarEq}), can be rewritten in
2057: terms of the conformal variables as
2058: \begin{eqnarray}
2059: \label{inflatonEqChao4}
2060: &&\chi_c'' - \nabla^2\chi_c -\frac{a''}{a}\chi_c + 
2061: (\chi_c^2 + q\phi_c^2)\chi_c  = 0 \\
2062: \label{scalarEqChao4}
2063: &&\phi_c'' - \nabla^2\phi_c -\frac{a''}{a}\chi_c + 
2064: q\chi_c^2\phi_c = 0\,,
2065: \end{eqnarray}
2066: where the prime denotes derivative with respect to conformal time. Since
2067: the universe expands as radiation-like in these scenarios,
2068: $a(\tau) \sim \tau$, so the terms proportional to $a''/a$
2069: in Eqs.~(\ref{inflatonEqChao4}) and (\ref{scalarEqChao4}) are soon zero, as
2070: explicitly checked in the simulations. Thanks to this, the
2071: model is conformal to Minkowski.
2072: 
2073: The parameter $q \equiv g^2/\lambda$ 
2074: controls the strength and width of the resonance.
2075: For the case of a massless inflaton without an extra scalar field,
2076: we just set $q = 0$ in Eq.~(\ref{inflatonEqChao4}) and
2077: ignore Eq.~(\ref{scalarEqChao4}). However, in that case, fluctuations
2078: of the inflaton also grow via parametric resonance. Actually, they
2079: grow as if they were fluctuations of a scalar field coupled to the
2080: zero-mode of the inflaton with effective couplig $q = g^2/\lambda =
2081: 3$, see Ref.~\cite{Greene}.
2082: 
2083: %%%%%%%%%%%%%%%%%%%%%%%%%
2084: \begin{figure}[htb]%[b]
2085: %\vspace{5mm}
2086: \begin{center}
2087: %\includegraphics[width=4.cm,height=7.cm,angle=-90]{meanFields.eps}
2088: \includegraphics[width=5.5cm,height=8.5cm,angle=-90]{spectraGW_chao4+Scalar.eps}
2089: \end{center}
2090: \vspace*{-5mm}
2091: \caption{The spectrum of the gravitational waves' energy density, 
2092: for coupled case with $\lambda = 10^{-14}$ and $g^2/\lambda = 120$. 
2093: The spectrum is shown accumulated up to different times during GW 
2094: production, so one can see its evolution. At each time, it is normalized 
2095: to the total instant density. This plot corresponds to a N = 128
2096: lattice simulation, from $\tau = 0$ to $\tau = 240$.}
2097: \label{fig6}
2098: \vspace*{-3mm}
2099: \end{figure}
2100: %%%%%%%%%%%%%%%%%%%%%%%%%
2101: 
2102: Following Refs.~\cite{TkachevGW} and~\cite{EastherLim}, we set
2103: $\lambda = 10^{-14}$ and $q$ = 120. Since, this case is also 
2104: computed in~\cite{DufauxGW}, we can also compare our results 
2105: with theirs. Moreover, we also present results for the pure 
2106: $\lambda\chi^4$ model with no extra scalar field, a case 
2107: only shown in Ref.~\cite{TkachevGW}.
2108: 
2109: We begin our simulations at the
2110: end of inflation, when the homogeneous inflaton verifies $\chi_0
2111: \approx 0.342 M_p$ and $\dot\chi_0 \approx 0$. We took initial quantum
2112: (conformal) fluctuations $1/\sqrt{2k}$ for all the modes up to a
2113: certain cut-off, and only added an initial zero-mode for the inflaton,
2114: $\chi_c(0) = 1$, $\chi_c(0)' = 0$. In Figs.~\ref{fig6} and~\ref{fig7}, 
2115: we show the 
2116: evolution of $\Omega_{_{GW}}$ during reheating, normalized to the 
2117: instant density at each time step, for the coupled and the pure 
2118: case, respectively. In the case with an extra scalar field, 
2119: the amplitude of the GWB saturates at the end of parametric 
2120: resonance, when the fields 
2121: variances have been stabilized. This is the beginnig of the turbulent 
2122: stage in the scalar fields, which seems not to source anymore the 
2123: production of GWs, as already stated in Refs.\cite{EastherLim,DufauxGW}.
2124: For the pure case, we also see the saturation of the amplitude of the 
2125: spectra, see Fig.~\ref{fig7}, although the long momenta tail seems to 
2126: slightly move toward higher values.
2127: 
2128: %%%%%%%%%%%%%%%%%%%%%%%%%
2129: \begin{figure}[htb]%[t]
2130: %\vspace{5mm}
2131: \begin{center}
2132: %\includegraphics[width=4.cm,height=7.cm,angle=-90]{meanFields.eps}
2133: \includegraphics[width=5.5cm,height=8.5cm,angle=-90]{spectraGW_chao4Pure.eps}
2134: \end{center}
2135: \vspace*{-5mm}
2136: \caption{The spectrum of the gravitational waves' energy density, 
2137: for the pure case, with $\lambda = 10^{-14}$. Again, we show 
2138: the spectrum accumulated up to different times during GW 
2139: production, normalized to the total instant density at each time.
2140: The plot corresponds to a N = 128 lattice simulation, 
2141: from $\tau = 0$ to $\tau = 2000$.}
2142: \label{fig7}
2143: \vspace*{-3mm}
2144: \end{figure}
2145: %%%%%%%%%%%%%%%%%%%%%%%%%
2146: 
2147: %%%%%%%%%%%%%%%%%%%%%%%%%
2148: \begin{figure}[htb]%[t]
2149: %\vspace{5mm}
2150: \begin{center}
2151: %\includegraphics[width=4.cm,height=7.cm,angle=-90]{meanFields.eps}
2152: \includegraphics[width=5.5cm,height=8.5cm,angle=-90]{spectraGW_chao4Today.eps}
2153: \end{center}
2154: \vspace*{-5mm}
2155: \caption{Today's ratio of gravitational waves normalized to radiation
2156: energy density, for both the coupled and the pure case. We took 
2157: $g_{*}/g_0 = 100$ to redshift the spectra from the time of the end 
2158: of production till today.}
2159: \label{fig8}
2160: \vspace*{-3mm}
2161: \end{figure}
2162: %%%%%%%%%%%%%%%%%%%%%%%%%
2163: 
2164: Of course, in either case, with and without
2165: an extra field $\phi$, in order to predict today's spectral window of
2166: the GW spectrum, we have, first, to normalize their energy density
2167: at the end of GW production to the total
2168: energy density at that moment. Secondly, we have to redshift the GW
2169: spectra from that moment of reheating, taking into account that the
2170: rate of expansion have changed significantly since the end of
2171: inflation, see Eq.(\ref{redshift}). In particular, the shape and 
2172: amplitude of GW spectra for the case with the extra scalar field
2173: coupled to the inflaton with $q = 120$, see Fig.~\ref{fig8}, 
2174: seems to coincide 
2175: with the espectra shown in Ref.~\cite{DufauxGW}. 
2176: On the other hand, we also reproduce in Fig.~\ref{fig8} a similar
2177: spectra to the one shown in~\cite{TkachevGW}, for the case of the pure
2178: quartic model. % with $\lambda = 10^{-14}$.  
2179: Thanks to the tremendous gain in
2180: computer power, we were able to resolve the 'spiky' pattern
2181: of the spectrum with great resolution. For the first time,
2182: it is clearly observed the exponential tail for large frequencies, see
2183: Figs.~\ref{fig7}, \ref{fig8}, not shown in Ref.~\cite{TkachevGW}.  The most
2184: remarkable fact, is that we also confirm that the peak structure in
2185: the GW power spectrum, see Fig.~\ref{fig7}, remains clearly visible at
2186: times much later than the one at which those peaks have dissapeared in
2187: the scalar fields' power spectrum. So, as pointed out in
2188: Ref.~\cite{TkachevGW}, this characteristic feature distinguish this
2189: particular model from any other.
2190: 
2191: Let us emphasize that we have run the simulations till times 
2192: much greater than that of the end of the resonance stage, both for 
2193: the pure and the coupled case. The role of the
2194: turbulence period after preheating seems, therefore, not to be very 
2195: important, despite its long duration. Apparently, the \textit{no-go} 
2196: theorem about the suppresion of GW at turbulence, discussed in~\cite{DufauxGW}, 
2197: is fulfilled. In Refs.~\cite{CEWB,Dufaux} it was 
2198: pointed out that gauge couplings or trilinear interactions could be 
2199: responsible for a fast thermalization of the universe after inflation 
2200: (see also Ref.~\cite{FK}), but as long as this takes place after the end 
2201: of the resonace stage, in principle this should not affect the results 
2202: shown above.
2203: 
2204: 
2205: \begin{widetext}
2206: 
2207: 
2208: %%%%%%%%%%%%%%%%%%%%%%%%%
2209: \begin{figure}[h]
2210: %\vspace{-5mm}
2211: \begin{center}
2212: \includegraphics[width=10cm,height=16cm,angle=-90]{sensitivityII.ps}
2213: \end{center}
2214: \vspace*{-5mm}
2215: \caption{The sensitivity of the different gravitational wave
2216: experiments, present and future, compared with the possible stochastic
2217: backgrounds; we include the White Dwarf Binaries (WDB)~\cite{WDB} and
2218: chaotic preheating ($\lambda\phi^4$, coupled and pure) for comparison.
2219: Note the two well differentiated backgrounds from high-scale and
2220: low-scale hybrid inflation. The bound marked (?) is estimated from 
2221: ultra high frequency laser interferometers' 
2222: expectations~\cite{Nishizawa2007}.}
2223: \label{fig5}
2224: \vspace*{-3mm}
2225: \end{figure}
2226: %%%%%%%%%%%%%%%%%%%%%%%%%
2227: 
2228: \end{widetext}
2229: 
2230: 
2231: \section{Conclusions}
2232: 
2233: 
2234: To summarize, we have shown that hybrid models are very efficient
2235: generators of gravitational waves at preheating, in three well defined
2236: stages, first via the tachyonic growth of Higgs modes, whose gradients
2237: act as sources of gravity waves; then via the collisions of highly
2238: relativistic bubble-like structures with large amounts of energy
2239: density, and finally via the turbulent regime (although this effect
2240: does not seem to be very significant in the presence of scalar
2241: sources), which drives the system towards thermalization. These
2242: waves remain decoupled since the moment of their production, and thus
2243: the predicted amplitude and shape of the gravitational wave spectrum
2244: today can be used as a probe of the reheating period in the very early
2245: universe. The characteristic spectrum can be used to distinguish
2246: between this stochastic background and others, like those arising from
2247: NS-NS and BH-BH coalescence, which are decreasing with frequency, or
2248: those arising from inflation, that are flat~\cite{SKC}.
2249: \\
2250: 
2251: We have plotted in Fig.~\ref{fig5} the sensitivity of planned GW
2252: interferometers like LIGO, LISA and BBO, together with the present
2253: bounds from CMB anisotropies (GUT inflation), from Big Bang
2254: Nucleosynthesis (BBN) and from milisecond pulsars (ms pulsar).  Also
2255: shown are the expected stochastic backgrounds of chaotic inflation
2256: models like $\lambda\phi^4$, both coupled and pure, as well as
2257: the predicted background from two different hybrid inflation models, a
2258: high-scale model, with $v=10^{-3}M_P$ and $\lambda\sim g^2\sim0.1$,
2259: and a low-scale model, with $v=10^{-5}M_P$ and $\lambda\sim g^2\sim
2260: 10^{-14}$, corresponding to a rate of expansion $H\sim 100$ GeV. The
2261: high-scale hybrid model produces typically as much gravitational waves
2262: from preheating as the chaotic inflation models. The advantage of
2263: low-scale hybrid models of inflation is that the background produced
2264: is within reach of future GW detectors like BBO~\cite{BBO}. It is
2265: speculated that future high frequency laser interferometers could
2266: be sensitive to a GWB in the MHz region~\cite{Nishizawa2007}, although
2267: they are still far from the bound marked with an interrogation sign.
2268: 
2269: For a high-scale model of inflation, we may never see the predicted GW
2270: background coming from preheating, in spite of its large amplitude,
2271: because it appears at very high frequencies, where no detector has yet
2272: shown to be sufficiently sensitive. On the other hand, if inflation
2273: occured at low scales, even though we will never have a chance to
2274: detect the GW produced during inflation in the polarization
2275: anisotropies of the CMB, we do expect gravitational waves from
2276: preheating to contribute with an important background in sensitive
2277: detectors like BBO. The detection and characterization of such a GW
2278: background, coming from the complicated and mostly unknown epoch of
2279: rehating of the universe, may open a new window into the very early
2280: universe, while providing a new test on inflationary cosmology.
2281: \\
2282: 
2283: \section*{Acknowledgments} 
2284: We wish to thank Andr\'es D\'iaz-Gil, Jean-Francois Dufaux, Richard
2285: Easther, Gary Felder, Margarita Garc\'\i a P\'erez, John T. Giblin
2286: Jr., Seiji Kawamura, Lev Kofman, Andrei Linde, Eugene A. Lim, Luis
2287: Fernando Mu\~noz-Mej\'{\i}as and Mischa Sall\'e for very useful
2288: comments, suggestions and constructive criticism. This work is
2289: supported in part by CICYT projects FPA2003-03801 and FPA2006-05807,
2290: by EU network "UniverseNet" MRTN-CT-2006-035863 and by CAM project
2291: HEPHACOS S-0505/ESP-0346. D.G.F. and A.S. acknowledges support from a
2292: FPU-Fellowship from the Spanish M.E.C.  We also acknowledge use of the
2293: MareNostrum Supercomputer under project AECT-2007-1-0005.
2294: 
2295: 
2296: 
2297: \begin{thebibliography}{99}
2298: 
2299: \bibitem{HulseTaylor}
2300: R.~A.~Hulse and J.~H.~Taylor,
2301:   %``Discovery of a pulsar in a binary system,''
2302:   Astrophys.\ J.\  {\bf 195}, L51 (1975);
2303:   %%CITATION = ASJOA,195,L51;%%
2304: Nobel Prize in Physics 1993,
2305:   {\tt http://nobelprize.org/nobel${}_{-}$\!prizes/physics/\\
2306:   laureates/1993/}
2307: 
2308: \bibitem{Maggiore}
2309: M.~Maggiore,
2310:   %``Gravitational wave experiments and early universe cosmology,''
2311:   Phys.\ Rept.\  {\bf 331}, 283 (2000);
2312:   %%CITATION = GR-QC 9909001;%%
2313: C.~J.~Hogan,
2314:   %``Gravitational wave sources from new physics,''
2315:   arXiv:astro-ph/0608567;
2316:   %%CITATION = ASTRO-PH 0608567;%%
2317: A.~Buonanno,
2318:   ``Gravitational waves,''
2319:   arXiv:0709.4682 [gr-qc].
2320:   %%CITATION = ARXIV:0709.4682;%%
2321: 
2322: \bibitem{Starobinsky} 
2323: A.~A.~Starobinsky,
2324:   %``Spectrum of relict gravitational radiation and the early state of the
2325:   %universe,''
2326:   JETP Lett.\  {\bf 30}, 682 (1979)
2327:   [Pisma Zh.\ Eksp.\ Teor.\ Fiz.\  {\bf 30}, 719 (1979)].
2328:   %%CITATION = ZFPRA,30,719;%%
2329: 
2330: \bibitem{LIGO}
2331: B.~Abbott {\it et al.}  [LIGO Scientific Collaboration],
2332:   %``Searching for a stochastic background of gravitational waves with LIGO,''
2333:   Astrophys.\ J.\  {\bf 659}, 918 (2007)
2334:   [arXiv:astro-ph/0608606].
2335:   %%CITATION = ASJOA,659,918;%%
2336:   LIGO Home Page: {\tt http://www.ligo.caltech.edu/}
2337: 
2338: \bibitem{LISA}
2339: S.~A.~Hughes,
2340:   %``LISA sources and science,''
2341:   arXiv:0711.0188 [gr-qc].
2342:   %%CITATION = ARXIV:0711.0188;%%
2343:   LISA Home Page: {\tt http://lisa.esa.int}
2344: 
2345: 
2346: \bibitem{BBO}
2347: V.~Corbin and N.~J.~Cornish,
2348:   %``Detecting the cosmic gravitational wave background with the big bang observer,''
2349:   Class.\ Quant.\ Grav.\  {\bf 23}, 2435 (2006);
2350:   %%CITATION = GR-QC 0512039;%%
2351: G.~M.~Harry, P.~Fritschel, D.~A.~Shaddock, W.~Folkner and E.~S.~Phinney,
2352:   %``Laser Interferometry For The Big Bang Observer,''
2353:   Class.\ Quant.\ Grav.\  {\bf 23}, 4887 (2006).
2354:   %%CITATION = CQGRD,23,4887;%%
2355:   BBO Home Page: {\tt http://universe.nasa.gov/new/program/bbo.html}
2356: 
2357: 
2358: \bibitem{DECIGO}
2359: S.~Kawamura {\it et al.},
2360:   %``The Japanese space gravitational wave antenna DECIGO,''
2361:   Class.\ Quant.\ Grav.\  {\bf 23}, S125 (2006).
2362:   %%CITATION = CQGRD,23,S125;%%
2363: 
2364: \bibitem{Planck}
2365: M.~Kamionkowski, A.~Kosowsky and A.~Stebbins,
2366:   %``A probe of primordial gravity waves and vorticity,''
2367:   Phys.\ Rev.\ Lett.\  {\bf 78}, 2058 (1997);
2368:   %%CITATION = ASTRO-PH 9609132;%%
2369: U.~Seljak and M.~Zaldarriaga,
2370:   %``Signature of gravity waves in polarization of the microwave background,''
2371:   Phys.\ Rev.\ Lett.\  {\bf 78}, 2054 (1997);
2372:   %%CITATION = ASTRO-PH 9609169;%%
2373: Planck Home page: {\tt http://www.rssd.esa.int/Planck}
2374: 
2375: \bibitem{BBN}
2376: B.~Allen, gr-qc/9604033.
2377:   %%CITATION = GR-QC 9604033;%%
2378: 
2379: \bibitem{pulsar}
2380: I.~H.~Stairs,
2381:   %``Testing General Relativity with Pulsar Timing,''
2382:   Living Rev.\ Rel.\  {\bf 6}, 5 (2003).
2383:   %%CITATION = ASTRO-PH 0307536;%%
2384: 
2385: \bibitem{Elena}
2386: T.~L.~Smith, E.~Pierpaoli and M.~Kamionkowski,
2387: %``A new cosmic microwave background constraint to primordial gravitational waves,''
2388:   Phys.\ Rev.\ Lett.\  {\bf 97}, 021301 (2006).
2389:   %%CITATION = ASTRO-PH 0603144;%%
2390: 
2391: 
2392: \bibitem{KosowskyTurner}
2393: A.~Kosowsky, M.~S.~Turner and R.~Watkins,
2394:   %``Gravitational waves from first order cosmological phase transitions,''
2395:   Phys.\ Rev.\ Lett.\  {\bf 69}, 2026 (1992),
2396: %%CITATION = PRLTA,69,2026;%%  
2397:   %``Gravitational radiation from colliding vacuum bubbles,''
2398:   Phys.\ Rev.\ D {\bf 45}, 4514 (1992);
2399: A.~Kosowsky and M.~S.~Turner,
2400:   %``Gravitational Radiation From Colliding Vacuum Bubbles: Envelope
2401:   %Approximation To Many Bubble Collisions,''
2402:   Phys.\ Rev.\ D {\bf 47}, 4372 (1993);
2403:   %%CITATION = ASTRO-PH 9211004;%%
2404: M.~Kamionkowski, A.~Kosowsky and M.~S.~Turner,
2405:   %``Gravitational radiation from first order phase transitions,''
2406:   Phys.\ Rev.\ D {\bf 49}, 2837 (1994).
2407:   %%CITATION = ASTRO-PH 9310044;%%
2408: 
2409: \bibitem{Nicolis}
2410: A.~Nicolis,
2411:   %``Relic gravitational waves from colliding bubbles and cosmic turbulence,''
2412:   Class.\ Quant.\ Grav.\  {\bf 21}, L27 (2004);
2413:   %%CITATION = GR-QC 0303084;%%
2414: C.~Grojean and G.~Servant,
2415:   %``Gravitational waves from phase transitions at the electroweak scale and
2416:   %beyond,''
2417:   arXiv:hep-ph/0607107.
2418:   %%CITATION = HEP-PH 0607107;%%
2419: 
2420: 
2421: \bibitem{Turbulence}
2422: A.~Kosowsky, A.~Mack and T.~Kahniashvili,
2423:   %``Gravitational radiation from cosmological turbulence,''
2424:   Phys.\ Rev.\ D {\bf 66}, 024030 (2002);
2425:   %%CITATION = ASTRO-PH 0111483;%%
2426: A.~D.~Dolgov, D.~Grasso and A.~Nicolis,
2427:   %``Relic backgrounds of gravitational waves from cosmic turbulence,''
2428:   Phys.\ Rev.\ D {\bf 66}, 103505 (2002).
2429:   %%CITATION = ASTRO-PH 0206461;%%
2430: G.~Gogoberidze, T.~Kahniashvili and A.~Kosowsky,
2431:   ``The spectrum of gravitational radiation from primordial turbulence,''
2432:   arXiv:0705.1733 [astro-ph].
2433:   %%CITATION = ARXIV:0705.1733;%%
2434: 
2435: 
2436: \bibitem{Nishizawa2007}
2437:   A.~Nishizawa {\it et al.},
2438:   ``Laser-interferometric Detectors for Gravitational Wave Background at 100
2439:   MHz : Detector Design and Sensitivity,''
2440:   arXiv:0710.1944 [gr-qc];
2441:   %%CITATION = ARXIV:0710.1944;%%
2442: %\bibitem{Cruise2006}
2443:   A.~M.~Cruise and R.~M.~J.~Ingley,
2444:   %``A prototype gravitational wave detector for 100-MHz,''
2445:   Class.\ Quant.\ Grav.\  {\bf 23}, 6185 (2006).
2446:   %%CITATION = CQGRD,23,6185;%%
2447: 
2448: \bibitem{preheating}
2449: L.~Kofman, A.~D.~Linde and A.~A.~Starobinsky,
2450:   %``Reheating after inflation,''
2451:   Phys.\ Rev.\ Lett.\  {\bf 73}, 3195 (1994);
2452:   %%CITATION = HEP-TH 9405187;%%
2453: L.~Kofman, A.~D.~Linde and A.~A.~Starobinsky,
2454:   %``Towards the theory of reheating after inflation,''
2455:   Phys.\ Rev.\ D {\bf 56}, 3258 (1997).
2456:   %%CITATION = HEP-PH 9704452;%%
2457: 
2458: 
2459: \bibitem{Thorne}
2460: K. S. Thorne, ``\textit{Gravitational Radiation}'' 
2461: in ``300 Hundred Years of Gravitation", 330-458. 
2462: 
2463: 
2464: \bibitem{GBF}
2465: J.~Garcia-Bellido and D.~G.~Figueroa,
2466:   %``A stochastic background of gravitational waves from hybrid preheating,''
2467:   Phys.\ Rev.\ Lett.\  {\bf 98}, 061302 (2007)
2468:   [arXiv:astro-ph/0701014].
2469: 
2470: 
2471: \bibitem{LythRep}
2472: D.~H.~Lyth and A.~Riotto,
2473:   %``Particle physics models of inflation and the cosmological density
2474:   %perturbation,''
2475:   Phys.\ Rept.\  {\bf 314}, 1 (1999).
2476:   %%CITATION = HEP-PH 9807278;%%
2477: 
2478: 
2479: \bibitem{WMAP}
2480: D.~N.~Spergel {\it et al.}, astro-ph/0603449;
2481:   %%CITATION = ASTRO-PH 0603449;%%
2482: M.~Tegmark {\it et al.}, astro-ph/0608632;
2483:   %%CITATION = ASTRO-PH 0608632;%%
2484: W.~H.~Kinney, E.~W.~Kolb, A.~Melchiorri and A.~Riotto,
2485:   %``Inflation model constraints from the Wilkinson microwave anisotropy probe
2486:   %three-year data,''
2487:   Phys.\ Rev.\ D {\bf 74}, 023502 (2006).
2488:   %%CITATION = ASTRO-PH 0605338;%%
2489: H.~Peiris and R.~Easther,
2490:   %``Slow roll reconstruction: Constraints on inflation from the 3 year WMAP
2491:   %dataset,''
2492:   astro-ph/0609003.
2493:   %%CITATION = ASTRO-PH 0609003;%%
2494: 
2495: 
2496: \bibitem{TkachevGW}
2497: S.~Y.~Khlebnikov and I.~I.~Tkachev,
2498:   %``Relic gravitational waves produced after preheating,''
2499:   Phys.\ Rev.\ D {\bf 56}, 653 (1997).
2500:   %%CITATION = HEP-PH 9701423;%%
2501: 
2502: 
2503: \bibitem{JuanGW}
2504: J.~Garcia-Bellido,
2505:   %``Preheating the universe in hybrid inflation,''
2506:   arXiv:hep-ph/9804205.
2507:   %%CITATION = HEP-PH/9804205;%%
2508: 
2509: 
2510: \bibitem{EastherLim}
2511: R.~Easther and E.~A.~Lim,
2512:   %``Stochastic gravitational wave production after inflation,''
2513:   JCAP {\bf 0604}, 010 (2006).
2514:   %%CITATION = ASTRO-PH 0601617;%%
2515: 
2516: \bibitem{EastherLim2}
2517: R.~Easther, J.~T.~.~Giblin and E.~A.~Lim,
2518:   ``Gravitational Wave Production At The End Of Inflation,''
2519:   arXiv:astro-ph/0612294.
2520:   %%CITATION = ASTRO-PH/0612294;%%
2521: 
2522: %\cite{Dufaux:2007pt}
2523: \bibitem{DufauxGW}
2524:   J.~F.~Dufaux, A.~Bergman, G.~N.~Felder, L.~Kofman and J.~P.~Uzan,
2525:   %``Theory and Numerics of Gravitational Waves from Preheating after
2526:   %Inflation,''
2527:   arXiv:0707.0875 [astro-ph].
2528:   %%CITATION = ARXIV:0707.0875;%%
2529: 
2530: \bibitem{Picasso}
2531: F.~Pegoraro, L.~A.~Radicati, P.~Bernard and E.~Picasso,
2532:   %``Electromagnetic detector for gravitational waves.''
2533:   Phys.\ Lett.\ A {\bf 68}, 165 (1978);
2534: G.~Gemme, A.~Chincarini, R.~Parodi, P.~Bernard and E.~Picasso,
2535:   gr-qc/0112021;
2536:   %%CITATION = GR-QC 0112021;%%
2537: R.~Ballantini {\it et al.},
2538:   gr-qc/0502054.
2539:   %%CITATION = GR-QC 0502054;%%
2540: 
2541: 
2542: 
2543: \bibitem{hybrid}
2544: A.~D.~Linde,
2545:   %``Hybrid inflation,''
2546:   Phys.\ Rev.\ D {\bf 49}, 748 (1994);
2547:   %%CITATION = ASTRO-PH 9307002;%%
2548: J.~Garcia-Bellido and A.~D.~Linde,
2549:   %``Preheating in hybrid inflation,''
2550:   Phys.\ Rev.\ D {\bf 57}, 6075 (1998).
2551:   %%CITATION = HEP-PH 9711360;%%
2552: 
2553: 
2554: \bibitem{tachyonic}
2555: G.~N.~Felder, J.~Garcia-Bellido, P.~B.~Greene, L.~Kofman, A.~D.~Linde and I.~Tkachev,
2556:   %``Dynamics of symmetry breaking and tachyonic preheating,''
2557:   Phys.\ Rev.\ Lett.\  {\bf 87}, 011601 (2001);
2558:   %%CITATION = HEP-PH 0012142;%%
2559: G.~N.~Felder, L.~Kofman and A.~D.~Linde,
2560:   %``Tachyonic instability and dynamics of spontaneous symmetry breaking,''
2561:   Phys.\ Rev.\ D {\bf 64}, 123517 (2001).
2562:    %%CITATION = HEP-TH 0106179;%%
2563: 
2564: \bibitem{symmbreak}
2565: J.~Garcia-Bellido, M.~Garcia Perez and A.~Gonzalez-Arroyo,
2566:   %``Symmetry breaking and false vacuum decay after hybrid inflation,''
2567:   Phys.\ Rev.\ D {\bf 67}, 103501 (2003);
2568:   %%CITATION = HEP-PH 0208228;%%
2569: 
2570: \bibitem{ester}
2571: J.~Garcia-Bellido and E.~Ruiz Morales,
2572:   %``Particle production from symmetry breaking after inflation and leptogenesis,''
2573:   Phys.\ Lett.\ B {\bf 536}, 193 (2002).
2574:   %%CITATION = HEP-PH 0109230;%%
2575: 
2576: 
2577: \bibitem{CEWB}
2578: J.~Garcia-Bellido, D.~Y.~Grigoriev, A.~Kusenko and M.~E.~Shaposhnikov,
2579:   %``Non-equilibrium electroweak baryogenesis from preheating after  inflation,''
2580:   Phys.\ Rev.\ D {\bf 60}, 123504 (1999).
2581:   %%CITATION = HEP-PH 9902449;%%
2582: J.~Garcia-Bellido, M.~Garcia-Perez and A.~Gonzalez-Arroyo,
2583:   %``Chern-Simons production during preheating in hybrid inflation models,''
2584:   Phys.\ Rev.\ D {\bf 69}, 023504 (2004);
2585:   %%CITATION = HEP-PH 0304285;%%
2586: A.~Tranberg and J.~Smit,
2587:   %``Baryon asymmetry from electroweak tachyonic preheating,''
2588:   JHEP {\bf 0311}, 016 (2003).
2589:   %%CITATION = HEP-PH 0310342;%%
2590: 
2591: \bibitem{magnetic}
2592: A.~Diaz-Gil, J.~Garcia-Bellido, M.~Garcia Perez and A.~Gonzalez-Arroyo,
2593:   %``Magnetic field production after inflation,''
2594:   PoS {\bf LAT2005}, 242 (2006);
2595:   %[arXiv:hep-lat/0509094].
2596:   %%CITATION = HEP-LAT 0509094;%%
2597:   PoS {\bf LAT2007}, 052 (2007),
2598:   arXiv:0710.0580 [hep-lat].
2599:   %%CITATION = NONE,,;%%
2600: 
2601: \bibitem{BSC} MareNostrum Supercomputer Home Page: \\
2602: {\tt http://www.bsc.es/}
2603: 
2604: \bibitem{IFTcluster} IFT cluster Home Page: \\
2605: {\tt http://lattice.ft.uam.es/iftcluster/}
2606: 
2607: \bibitem{LatticeEasy}
2608: G. N. Felder and I. Tkachev, arXiv: hep-ph/0011159;\\
2609: {\tt http://www.science.smith.edu/departments/\\
2610: Physics/fstaff/gfelder/latticeeasy/}
2611: 
2612: \bibitem{GBLW}
2613: J.~Garcia-Bellido, A.~D.~Linde and D.~Wands,
2614:   %``Density perturbations and black hole formation in hybrid inflation,''
2615:   Phys.\ Rev.\  D {\bf 54}, 6040 (1996).
2616:   %%CITATION = PHRVA,D54,6040;%%
2617: 
2618: \bibitem{MSSMinf}
2619: R.~Allahverdi, K.~Enqvist, J.~Garcia-Bellido and A.~Mazumdar,
2620:   %``Gauge invariant MSSM inflaton,''
2621:   Phys.\ Rev.\ Lett.\  {\bf 97}, 191304 (2006);
2622:   %%CITATION = PRLTA,97,191304;%%
2623: R.~Allahverdi, K.~Enqvist, J.~Garcia-Bellido, A.~Jokinen and A.~Mazumdar,
2624:   JCAP {\bf 0706} (2007) 019 [arXiv:hep-ph/0610134].
2625:   %%CITATION = HEP-PH/0610134;%%
2626: 
2627: 
2628: 
2629: \bibitem{classical}
2630: S.~Y.~Khlebnikov and I.~I.~Tkachev,
2631:   %``Classical decay of inflaton,''
2632:   Phys.\ Rev.\ Lett.\  {\bf 77}, 219 (1996);
2633:   %%CITATION = HEP-PH 9603378;%%
2634: %``Resonant decay of Bose condensates,''  Phys.\ Rev.\ Lett.\  
2635: {\bf 79}, 1607 (1997);
2636:   %%CITATION = HEP-PH 9610477;%%
2637: T.~Prokopec and T.~G.~Roos,
2638:   %``Lattice study of classical inflaton decay,''
2639:   Phys.\ Rev.\ D {\bf 55}, 3768 (1997).
2640:   %%CITATION = HEP-PH 9610400;%%
2641: 
2642: 
2643: \bibitem{Mukhanov}
2644: V.~F.~Mukhanov, H.~A.~Feldman and R.~H.~Brandenberger,
2645:   %``Theory of cosmological perturbations. 
2646:   %Part 1. Classical perturbations. Part 2. Quantum theory of 
2647:   %perturbations. Part 3. Extensions,''
2648:   Phys.\ Rept.\  {\bf 215}, 203 (1992).
2649:   %%CITATION = PRPLC,215,203;%%
2650: 
2651: 
2652: \bibitem{Carroll}
2653: S.~Carroll, ``\textit{Spacetime and Geometry: An introduction to General
2654: Relativity},'' Addison Wesley (2003).
2655: 
2656: 
2657: \bibitem{Wilson}
2658: K.~G.~Wilson,
2659:   %``CONFINEMENT OF QUARKS,''
2660:   Phys.\ Rev.\  D {\bf 10}, 2445 (1974).
2661:   %%CITATION = PHRVA,D10,2445;%%
2662: 
2663: \bibitem{Cooper}
2664: F.~Cooper, S.~Habib, Y.~Kluger, E.~Mottola, J.~P.~Paz and P.~R.~Anderson,
2665:   %``Nonequilibrium quantum fields in the large N expansion,''
2666:   Phys.\ Rev.\  D {\bf 50}, 2848 (1994)
2667:   [arXiv:hep-ph/9405352];
2668:   %%CITATION = PHRVA,D50,2848;%%
2669: F.~Cooper, S.~Habib, Y.~Kluger and E.~Mottola,
2670:   %``Nonequilibrium dynamics of symmetry breaking in lambda Phi**4 field
2671:   %theory,''
2672:   Phys.\ Rev.\  D {\bf 55}, 6471 (1997)
2673:   [arXiv:hep-ph/9610345].
2674:   %%CITATION = PHRVA,D55,6471;%%
2675: 
2676: \bibitem{Vega}
2677: D.~Boyanovsky, D.~Cormier, H.~J.~de Vega, R.~Holman, A.~Singh and M.~Srednicki,
2678:   %``Scalar field dynamics in Friedman Robertson Walker spacetimes,''
2679:   Phys.\ Rev.\  D {\bf 56}, 1939 (1997)
2680:   [arXiv:hep-ph/9703327];
2681:   %%CITATION = PHRVA,D56,1939;%%
2682: D.~Boyanovsky, H.~J.~de Vega, R.~Holman and J.~Salgado,
2683:   %``Non-equilibrium Bose-Einstein condensates, dynamical scaling and  
2684:   %symmetric evolution in large N phi**4 theory,''
2685:   Phys.\ Rev.\  D {\bf 59}, 125009 (1999)
2686:   [arXiv:hep-ph/9811273].
2687:   %%CITATION = PHRVA,D59,125009;%%
2688: 
2689: \bibitem{Baacke}
2690: J.~Baacke, K.~Heitmann and C.~Patzold,
2691:   %``Nonequilibrium dynamics: A renormalized computation scheme,''
2692:   Phys.\ Rev.\  D {\bf 55}, 2320 (1997)
2693:   [arXiv:hep-th/9608006];
2694:   %%CITATION = PHRVA,D55,2320;%%
2695: Phys.\ Rev.\  D {\bf 55}, 7815 (1997)
2696:   [arXiv:hep-ph/9612264];
2697:   %%CITATION = PHRVA,D55,7815;%%
2698: %``Renormalization of nonequilibrium dynamics in FRW cosmology,''
2699:   Phys.\ Rev.\  D {\bf 56}, 6556 (1997)
2700:   [arXiv:hep-ph/9706274].
2701:   %%CITATION = PHRVA,D56,6556;%%
2702: 
2703: \bibitem{BBKS}
2704: J.~M.~Bardeen, J.~R.~Bond, N.~Kaiser and A.~S.~Szalay,
2705:   %``The Statistics Of Peaks Of Gaussian Random Fields,''
2706:   Astrophys.\ J.\  {\bf 304}, 15 (1986).
2707:   %%CITATION = ASJOA,304,15;%%
2708: 
2709: \bibitem{MichaTkachev}
2710: R.~Micha and I.~I.~Tkachev,
2711:   %``Relativistic turbulence: A long way from preheating to equilibrium,''
2712:   Phys.\ Rev.\ Lett.\  {\bf 90}, 121301 (2003);
2713:   %%CITATION = HEP-PH 0210202;%%
2714: %R.~Micha and I.~I.~Tkachev,
2715:   %``Turbulent thermalization,''
2716:   Phys.\ Rev.\ D {\bf 70}, 043538 (2004).
2717:   %%CITATION = HEP-PH 0403101;%%
2718: 
2719: \bibitem{LATTICEWEB}
2720: http://lattice.ft.uam.es/gw
2721: 
2722: \bibitem{Weinberg}
2723: S. Weinberg, ``\textit{Gravitation and Cosmology}'', 
2724: John Wiley $\&$ Sons, 1972.
2725: 
2726: 
2727: \bibitem{Greene}
2728: P.~B.~Greene, L.~Kofman, A.~D.~Linde and A.~A.~Starobinsky,
2729:   %``Structure of resonance in preheating after inflation,''
2730:   Phys.\ Rev.\  D {\bf 56}, 6175 (1997)
2731:   [arXiv:hep-ph/9705347].
2732:   %%CITATION = PHRVA,D56,6175;%%
2733: 
2734: 
2735: \bibitem{WDB}
2736: A.~J.~Farmer and E.~S.~Phinney,
2737:   %``The gravitational wave background from cosmological compact binaries,''
2738:   Mon.\ Not.\ Roy.\ Astron.\ Soc.\  {\bf 346}, 1197 (2003).
2739:   %%CITATION = ASTRO-PH 0304393;%%
2740: 
2741: 
2742: \bibitem{Dufaux}
2743: J.~F.~Dufaux, G.~N.~Felder, L.~Kofman, M.~Peloso and D.~Podolsky,
2744:   %``Preheating with trilinear interactions: Tachyonic resonance,''
2745:   JCAP {\bf 0607}, 006 (2006).
2746:   %%CITATION = JCAPA,0607,006;%%
2747: 
2748: \bibitem{FK}
2749: G.~N.~Felder and L.~Kofman,
2750:   %``The development of equilibrium after preheating,''
2751:   Phys.\ Rev.\  D {\bf 63}, 103503 (2001).
2752:   %%CITATION = PHRVA,D63,103503;%%
2753: 
2754: \bibitem{SKC}
2755: T.~L.~Smith, M.~Kamionkowski and A.~Cooray,
2756:   %``Direct detection of the ifnlationary gravitational wave background,''
2757:   Phys.\ Rev.\ D {\bf 73}, 023504 (2006).
2758:   %%CITATION = ASTRO-PH 0506422;%%
2759: 
2760: 
2761: \end{thebibliography}
2762: 
2763: \end{document}
2764: 
2765: